Vous êtes sur la page 1sur 33

Palaeontologia Electronica

http://palaeo-electronica.org

MICROFACIES OF STROMATOLITIC SINTER FROM


ACID-SULPHATE-CHLORIDE SPRINGS AT PARARIKI STREAM,
ROTOKAWA GEOTHERMAL FIELD, NEW ZEALAND
Richard Schinteie, Kathleen A. Campbell, and Patrick R.L. Browne
ABSTRACT
We present a unique, scale-integrated, and spatially controlled study of acidderived sinters and their abiotic-biotic relations. Through a microfacies-based
approach, we provide context and constraints for inferring causal factors in the formation of these sinters. Four distinct microfacies of siliceous stromatolitic sinter formation
and their associated microbiota were elucidated from acid-sulphate-chloride hot spring
outflows (pH 2.1-2.3, 91-30C), located on the floodplain of Parariki Stream, ~1 km
north of Lake Rotokawa in the Rotokawa Geothermal Field. Microfacies 1 comprises
cup- to ridge-shaped sinters forming close to vents (91-64C) with relatively high water
and gas discharge. Sinter surfaces are characterised by relatively small (0.5 cm high)
spicules, irregular, gnarly siliceous textures and colonisation by coccoidal microorganisms (1-1.5 m in diameter). Microfacies 2 consists of spiculose (1 cm high) sinters colonised by bacilli (1-2.3 m long), diatoms and coccoidal algae (210 m in diameter)
that are surrounded by quiescent waters (85-30C) with little steam discharge. Microfacies 3 is typified by parallel-laminated sinters forming on slightly steepened areas that
are colonised by bacilli (1-8 m long), diatoms and coccoidal algae (210 m in diameter) and exposed to fluctuating water levels (60-54C). Microfacies 4 constitutes thin
siliceous sinter rims forming mainly on small pumiceous clasts that rest upon moist
(67-45C) sandy substrates and colonised by bacilli (1-2.3 m long), diatoms and
spherical cells (2-6 m in diameter). Sinter morphology, texture and formation mechanisms, as well as microbial colonisation, depend on a variety of environmental constraints that can act at a scale of centimetres or less. Textural development of the
sinters, including their laminae, is attributed to a combination of abiotic and biotic factors. The differential preservation potentials of microbial communities need to be taken
into account when assessing biodiversity of ancient sinters.
Richard Schinteie. Geology Programme, School of Geography, Geology and Environmental Science,
University of Auckland, Private Bag 92019, Auckland 1142, New Zealand. Currently: Research School of
Earth Sciences, Building 61, Mills Road, The Australian National University, Canberra A.C.T 0200,
Australia.
richard.schinteie@anu.edu.au

PE Article Number: 10.1.4A


Copyright: Paleontological Society April 2007
Submission: 5 September 2006. Acceptance: 24 January 2007
Schinteie, Richard, Campbell, Kathleen A., and Browne, Patrick R.L., 2007. Microfacies of Stromatolitic Sinter from Acid-sulphatechloride Springs at Parariki Stream, Rotokawa Geothermal Field, New Zealand. Palaeontologia Electronica Vol. 10, Issue 1; 4A:33p,
8.8MB;
http://palaeo-electronica.org/paleo/2007_1/sinter/index.html

SCHINTEIE, CAMPBELL, & BROWNE: STROMATOLITIC MICROFACIES

Kathleen A. Campbell. Geology Programme, School of Geography, Geology and Environmental Science,
University of Auckland, Private Bag 92019, Auckland 1142, New Zealand.
ka.campbell@auckland.ac.nz (corresponding author)
Patrick R.L. Browne. Geology Programme, School of Geography, Geology and Environmental Science,
University of Auckland, Private Bag 92019, Auckland 1142, New Zealand.
prl.browne@auckland.ac.nz

KEY WORDS: sinter, silica, stromatolite, microorganisms, geothermal, Rotokawa, New Zealand
INTRODUCTION
Stromatolites may serve as important proxies
for early life on Earth. Many modern siliceous hot
spring deposits, or sinters, have laminated growth
structures that characterise such stromatolites
(e.g., Walter et al. 1972; Doemel and Brock 1974;
Renaut et al. 1998; Jones et al. 2000, 2005; Konhauser et al. 2001; Campbell et al. 2002; Guidry
and Chafetz 2003; Handley et al. 2005). Sinters
are often colonised by a range of (hyper)thermophilic microbiota that can become silicified and
incorporated into these deposits (e.g., Inagaki et al.
2001; Blank et al. 2002; Walker et al. 2005). The
taxonomic identities of these microbes can vary
within and between different geothermal settings,
and are a result of the numerous niches and
(micro)habitats encountered in these localities
(e.g., Schinteie 2005; Pancost et al. 2005, 2006).
Microbes inhabiting thermal environments are
often placed at the most deeply rooted parts of the
universal tree of life (e.g., Barns et al. 1996; Stetter
1996; Pace 1997; Hugenholtz et al. 1998). Hence,
the mineralisation of sinters and the phylogeny and
ecology of hot spring microorganisms are central
themes in studies concerning the origin of life,
astrobiology and mineral-microbe interactions
(e.g., Walter and Des Marais 1993; Henley 1996;
Farmer and Des Marais 1999; Farmer 2000; Handley et al. 2005). Indeed, hot spring deposits, like
sinters, may have provided surfaces that concentrated organic chemical constituents, thereby contributing to the formation of biological membranes
important for the origin of life (Henley 1996).
Studies concerned with the formation of sinter
and the distribution of their associated microbiota
require an understanding of the different litho- and
biofacies occurring at hot spring sites (Farmer
2000). The deposition of actively forming sinters
and the occurrence of specific microorganisms are
affected by shifting environmental parameters that
result in zonations around hot spring effluents with
respect to sinter morphology, texture, and microbial
species composition. In particular, deposits of sin2

ter and their microbiota can act as sensitive indicators of pH-temperature conditions (e.g., Brock
1978; Cassie and Cooper 1989; Cady and Farmer
1996; Lowe et al. 2001; Jones et al. 2000; Jones
and Renaut 2003; Lynne and Campbell 2003,
Rodgers et al. 2004) and the hydrodynamics of
spring discharge (e.g., Jones and Renaut 1997;
Braunstein and Lowe 2001; Lowe et al. 2001;
Guidry and Chafetz 2003; Lowe and Braunstein
2003). From observed variations in the texture and
distribution of sinters, and the species composition
or cell morphology of the associated microbiota,
facies models have been constructed to characterise surficial deposits from alkali-chloride hot
springs (e.g., Walter 1976; Cady and Farmer 1996;
Farmer 2000; Campbell et al. 2001; Guidry and
Chafetz 2003; Lowe and Braunstein 2003).
While ancient sinters from extinct hot springs
may retain some primary textural characteristics
(e.g., Rice et al. 1995; Walter et al. 1996; Trewin et
al. 2003; Campbell et al. 2001, 2004), their palaeoenvironmental
signatures
commonly
are
obscured by the loss of fine-scale microstructure
due to diagenetic overprinting (Cady and Farmer
1996) or differential preservation (Guidry and
Chafetz 2003). In addition, interpreting ancient sinter facies requires an understanding of the relative
contributions of abiotic and biotic factors in the formation of a variety of sinter textures. However,
studies of modern, actively forming sinters allow for
such shortcomings in palaeoenvironmental reconstruction to be addressed (Farmer 1999).
While most investigations are of sinters
deposited from thermal waters of near neutral pH,
very few (e.g., Jones et al. 1999, 2000; Mountain et
al. 2003; Rodgers et al. 2004) have addressed sinters formed from highly acidic (pH ~3 or lower) hot
spring waters. Such studies could potentially
enable the recognition of extinct hot spring systems and their deposits, even where there is no
longer any evidence of the original discharging
waters. Indeed, acid hot spring deposits may be
the most appropriate terrestrial analogues for the

PALAEO-ELECTRONICA.ORG

recognition of extraterrestrial hydrothermal deposits. Acidic hydrothermal fluids and solphatara-like


environments may occur on Mars (e.g., Farmer
1996, 2000; McCollom and Hynek 2005). In addition, numerous landforms and sedimentary features observed on Mars appear to have
equivalents in acid environments on Earth (Benison and Laclair 2002; Laclair and Benison 2002;
Bullock 2005). Chemical and mineralogical data
obtained from previous Mars missions also paint a
picture of acid alteration (e.g., Benison and Laclair
2002; Kerr 2004; Bullock 2005; McCollom and
Hynek 2005).
This study aims to further contribute to the
understanding of acid-derived sinters. We investigated microfacies (i.e., facies changes over centimetres or less) of siliceous stromatolitic sinters
formed in acid-sulphate-chloride spring outflows
(91-30C, pH 2.1-2.3) located on the floodplain of
Parariki Stream, Rotokawa Geothermal Field, New
Zealand. This field provides a rare setting where
acid fluids rich in silica deposit sinter. The deposits
follow strong environmental gradients that result in
distinctive morphological, textural, and microbial
characteristics. Molecular, DNA-based surveys at
the site have shown that microbial communities
vary markedly with respect to species composition
between different microfacies (Schinteie 2005).
Sequences of 16S ribosomal RNA (rRNA) genes
were extracted from individual sinters and were
related to numerous archaeal, bacterial, and
eukaryotic thermoacidophiles. The results of this
molecular survey are the subject of a different contribution that closely ties in with the results outlined
herein.
We employed an integrative approach with
the application of: (1) detailed mapping of sinter
occurrences with respect to temperature, hydrodynamics of spring discharge, and sinter formation
rates; (2) X-ray powder diffractometry (XRPD) to
characterise sinter mineralogy; (3) thermal analysis
for investigating water content of the sinters; (4)
thin-section petrography for recognising broad sinter textures; and (5) Scanning Electron Microscopy
(SEM) to resolve textural micro-structures and mineral-microbe relationships. Treatment of fresh sinter samples with glutaraldehyde slowed the
deterioration of associated microbial communities,
lessened the risk of contamination by post-collection microbial overgrowth, and better illuminated
the role of microorganisms in acid-derived sinter
texture formation (cf. Cady and Farmer 1996;
Lynne and Campbell 2003; Handley et al. 2005).

MATERIALS AND METHODS


In 2004, the study site was mapped to determine sinter microfacies types, vent locations, fluid
flow directions, temperatures, and pHs of discharged thermal waters, water level changes, and
silica accretion rates. The site was visited several
times that year to note any environmental changes,
especially of water levels. Mapping was undertaken with tape and compass. Temperatures and
pHs were measured with a portable battery operated OrionTM (model 250A) pH/ISE meter with
automatic temperature compensation. Silica accretion rates were measured by placing glass slides
(25 x 75 mm) vertically in the discharge channels
for 67 days (cf. Mountain et al. 2003; Handley
2004; Handley et al. 2005). Filtered (<0.2 m)
water samples were collected in February 2006 to
determine anion (unacidified) and cation (acidified
by 0.5% HNO3) concentrations. For HCO3 and
H2S concentrations, samples were collected in rubber-sealed bottles. The waters were analysed by
the Institute of Geologic and Nuclear Sciences
(IGNS), New Zealand.
The silica mineralogy of the sinters was determined by X-ray powder diffractometry (XRPD).
Each sample was air-dried and ground to a fine
powder in a mortar and pestle. XRPD was conducted with a Philips diffraction goniometer fitted
with a graphite monochromator. Samples were
scanned at 1.95 2/min with a step size of 0.02
from 2 to 62 2. Operating conditions were 40 kV
and 20 mA using CuK radiation (1 = 1.54051
; 2 = 1.5443 ).
Sinter composition and water content was
determined by combined differential thermal analysis (DTA) and thermogravimetric analysis (TGA).
Sinters were air-dried, ground to a fine powder,
placed into pre-weighed platinum crucibles (~0.22
g), and heated to 1400C in a Polymer Laboratories Simultaneous Thermal Analyser 1500 (Epsom,
Surrey, UK) equipped with PLus V software
(v.5.40) and a type R (Pt-13% Rh/Pt) flat-plate thermocouple system. Heating was in dry air or nitrogen. The former allowed the combustion of organic
material to be gauged, while the latter measured
water loss.
Fine-scale sinter textures and potential microbial involvements were examined by Scanning
Electron Microscopy (SEM). Samples were stored
in a fixative of 2.5% glutaraldehyde immediately
upon field collection. Prior to imaging, samples
were critical-point dried to prevent the effects of
surface tension from destroying delicate biofilms.
3

SCHINTEIE, CAMPBELL, & BROWNE: STROMATOLITIC MICROFACIES

178E

174E

2
176E

177E

36S

40S

38S

Lake
Rotorua

o
Waikat

Parariki Stream

Rive
r

N
Waikato River

Rotokawa

Study site

Lake
Taupo

Lake Rotokawa

Resistivity boundary

Geothermal field

Deep drillhole

Active volcano
Lake

Chloride spring

39S

River
TVZ boundary

Area of steaming ground


and acid-sulphate-chloride springs
Alluvium
Hydrothermal eruption breccia

30 km

1 km

Taupo Pumice

Figure 1. Study site location. (1) The Taupo Volcanic Zone and associated geothermal fields in the central North
Island, New Zealand. Modified from Bibby et al. (1995) and Wilson et al. (1995). (2) Map of the Rotokawa Geothermal Field and location of the study site. Modified from Collar and Browne (1985), Krupp and Seward (1987) and
Reyes et al. (2002).

The samples were rinsed twice in deionised water


followed by dehydration through an ethanol series
(30%, 50%, 70%, 90%, and 100% x 2). Ethanol
was exchanged for liquid CO2 at 5500 kPa by
flushing and soaking for one hour in a Polaron
E3000 Series II critical point drier (Polaron Equipment Ltd, Watford, UK). Liquid CO2 was subsequently vaporised at 31.5 C. Critical-point dried
samples were mounted on aluminium stubs and
coated with platinum using a Polaron SC 7640
Sputter Coater (Quorum Technologies Ltd,
Newhaven, UK). Samples were examined with a
Philips SEM XL30S (Eindhoven, Netherlands) fitted with SiLi (lithium drifted) electron-dispersive Xray spectrometer (EDS) with a Supra Ultra Thin
Window (EDAX, Mahwah, New Jersey, USA) for
assessing elemental composition of sinter components.
To examine broad sinter textures and allow for
comparisons with fine-scale SEM images, thin sections of the sinters were constructed. Samples
were first embedded overnight in an epoxy resin of
methyl methacrylate, and the dried resin was subsequently polymerised using cobalt-60 radiation.
Sections were examined under a Nikon Labophot4

POL petrographic microscope (Tokyo, Japan)


equipped with a Nikon Coolpix 4500 digital camera
(Tokyo, Japan).
SETTING AND SITE DESCRIPTION
The Rotokawa geothermal field is situated in
the southeastern margin of the Taupo Volcanic
Zone, North Island, New Zealand (Figure 1.1). The
field covers an area of ~6 km2 that is characterised
by steaming ground, fumaroles, hot springs, mud
pots, collapse pits, and infilled hydrothermal eruption craters (Figure 1.2) (Browne 1974; Krupp et al.
1986; Krupp and Seward 1987). Holocene pumice
of the Taupo Subgroup and hydrothermal eruption
breccia cover much of the surface of the thermal
area (Healy 1965; Vucetich and Pullar 1973; Collar
and Browne 1985). Hot springs discharge predominantly acid-sulphate-chloride waters (pH of 2-4)
(Ellis and Wilson 1961; Jones et al. 2000; Teece
2000; Mountain et al. 2003). The most prominent
surface features are the acidic (pH ~2.3) Lake
Rotokawa of 0.6 km2 and the main thermal area
directly to the north (Forsyth 1977; Krupp et al.
1986; Krupp and Seward 1987). On the northeast-

PALAEO-ELECTRONICA.ORG

Figure 2. Overview of the study site with view to the


south. Parariki Stream to the left is 2 m wide and flowing
north.

ern margin of Lake Rotokawa, hot spring activity


occurs in a terrace called Sinter Flat (Krupp et al.
1986; Krupp and Seward 1987). Several thin siliceous deposits and stromatolitic growth features
have been described from this terrace (e.g., Krupp
et al. 1986; Krupp and Seward 1987; Jones et al.
2000; Teece 2000; Mountain et al. 2003).
Lake Rotokawa drains via Parariki Stream in a
northeast direction from the east side of the lake
into the Waikato River (Figure 1.2). Numerous hot
springs situated on the stream banks discharge silica-rich, acid-sulphate-chloride waters along the
southern part of the stream (Grange 1937; Forsyth
1977; Krupp and Seward 1987; Teece 2000).
The study site is situated on a pumiceous
floodplain (Figure 2), located primarily alongside
the western bank of the Parariki Stream, ~1 km
northeast of Lake Rotokawa (Figure 1.2, red
arrow). This floodplain is ~30 m long, up to 7 m

wide and separated from the stream by a series of


riffles, composed of well sorted coarse pumiceous
alluvium (Figure 3). Through most of the year in
2004, the stream was confined to its channel and
did not infiltrate the floodplain. During summer,
thermal water levels were often lower than during
winter. Hot spring vents at this site discharge
mostly clear acid-sulphate-chloride waters (Table
1) ranging from 91 to 40C with an average pH of
~2.2 (Figure 4). Teece (2000) determined that silica
in the Parariki discharge waters is predominantly
monomeric. The spring water discharging into the
Parariki Stream and the study site contain about
300 mg kg-1 SiO2 (Table 1). They are thus very
close to saturation with respect to amorphous silica
but oversaturated in silica with respect to opal-CT,
opal-C and quartz (cf. Fournier, 1985; Dove and
Rimstidt, 1994). The relatively high silica contents
are a legacy of the >320C equilibration temperatures in the geothermal reservoir (Krupp et al.
1986; Krupp and Seward 1987), with silica derived
from depth, rather than being leached from the surface rocks, as is the case in many acid-sulphate
areas (cf. White et al. 1956; Rodgers et al. 2002,
2004). For this reason, deposits forming directly
from these silica-rich waters are interpreted to be
silica sinter, rather than silica residue (Rodgers et
al. 2004).
At the study site, four main types of siliceous
sinter deposits were discerned (Figure 3). They are
defined herein as four distinct sedimentary microfacies based on temperature and hydrodynamics of
surrounding thermal water, the geometry, size and
texture of the sinters, as well as their associated
microbiota.
MICROFACIES CHARACTERISTICS AND
SINTER PROPERTIES
Microfacies 1 Cup- to Ridge-Shaped Sinter
Microfacies 1 (Figure 3) sinter forms in slightly
turbulent waters with high emissions of steam, and
water temperatures ranging from 91C at some
vents to 64C in the associated discharge channels
(Figure 4.1). The water turbulence is caused by the

Table 1 Composition, temperature, and pH values of representative sampling localities at 11 and 21 (see Figure 4).
Concentrations are in mg kg-1.
Location

T (C)

pH

Li

Na

Mg

Ca

Al

Fe

SiO2

As

Cl

HCO3

H 2S

SO4

11

72

2.1

2.8

333

31

33

17.1

39

15.7

297

0.63

448

25

0.47

1340

21

84.6

2.26

2.5

313

32

6.9

34

15.5

28

15.8

302

0.51

393

26

3.4

1282

SCHINTEIE, CAMPBELL, & BROWNE: STROMATOLITIC MICROFACIES

N
Key:
Microfacies 1 sinter (substrate width >2 cm)
Microfacies 1 sinter (substrate width <2 cm)
Microfacies 2 sinter (substrate width >2 cm)
Microfacies 2 sinter (substrate width <2 cm)

Microfacies 3 sinter
Microfacies 4 sinter
Dry cracked siliceous sinter
(indicating former thermal discharge)

Sulphur accumulation
Brown vent deposit
Alluvial mud
Dry alluvial sand
Submerged alluvium
Pumice clasts
Silicified pumiceous tuff
Vegetation (manuka)
Thermal water flow

Stream water flow

Scale:
0

5m

Figure 3. Map of the study site showing the spatial distribution of the four distinct sedimentary microfacies of siliceous
sinter formation. Also shown are thermal and stream water flow directions, the distribution of alluvium, sulphur accumulations, outcrops of silicified pumice tuff, and vegetation.
6

PALAEO-ELECTRONICA.ORG

ebullient discharge from nearby vents that produces pulses of thermal water (typically 85C)
washing onto the siliceous sinters. Neither splash
nor spray was observed around these deposits.
However, water level changes were recorded in
this microfacies (Figure 4.1). Sinter forms subaerially on pumice clasts and pine cones that act as
substrates above water level (Figure 5). Sinter
accretion rates were fastest at this microfacies. An
average of 0.24 g silica accumulated on glass
slides over 67 days in this environment (Figure 4).
It is unknown if Microfacies 1 sinter accreted at a
uniform rate or not.
XRPD analyses of sinters from this and all
other microfacies of this study site showed that
they are composed almost entirely of opal-A, with
minor traces of detrital quartz and/or feldspar (for
full width at half maximum intensity (FWHM) values
and density/porosity data, see Rodgers et al.
2004). DTA of sinters from this study also gave
identical traces characteristic of opal-A, with no
indications of clay content (cf. Herdianita et al.
2000a). TGA of these sinters indicated weight loss
of approximately 5 to 6% resulting from opal-A
dehydration (cf. Herdianita et al. 2000b; Handley et
al. 2005).
Microfacies 1 sinter is vitreous, light grey in
colour, and usually cup- or ridge-shaped with
minute microspicules (0.5 cm high) visible on the
uppermost sides of the rim or within cavities (Figure 5.2). The sinter surfaces have multiple, thin layers of silica with an irregular texture, and are
composed of gnarled, broken, and isolated surface
remnants (Figure 5.4). This irregular texture is gradational, occurring predominantly in the lowermost
portions of the sinter, close to the air-water interface. Vertical sections through the sinters exhibit
alternating series of light brown and light grey laminae (Figure 5.5). The laminae range from <2 to 50
m in thickness and alternate between brown to
dark brown and clear, translucent silica (Figure
5.6). No evidence for the presence of microorganisms or any mineral-microbe association was
observed in thin-section. Laminae in the lower portions of the sinters are relatively flat, but become
progressively more convex upwards. On and within
these layers, conical microspicules may be
present, but in a relatively lower frequency than in
Microfacies 2 (see below). Under SEM, vertical
sections appeared vitreous and massive, without
any conspicuous laminae. Prokaryote-sized microorganisms were only rarely observed in these sections.

At the Parariki site, sinter morphology is


affected by the subaerially exposed substrate
dimensions. Substrates that are relatively wider
and higher above water exhibit limited siliceous
coating. This coating is confined to the outer
fringes of the substrates, resulting in the formation
of cup-shaped deposits (Figure 5.2). On substrates
that are low-lying and less aerially extensive, silica
deposited across the substrate and a ridge-shaped
deposit arises (Figure 5.3).
Large areas of Microfacies 1 sinter surfaces
are covered by irregularly lobed, coccoidal microorganisms (1-1.5 m in diameter) (Figure 6.1). These
microbes are linked to one another through a
meshwork of mucosal to filamentous extracellular
polymeric substances (EPS). Such microbe-EPS
assemblages are known as biofilms (cf. Cady and
Farmer 1996; Handley et al. 2005). Biofilms in the
upper- and inner-most portions of the Parariki sinters typically assumed the shapes of isolated, conical cell clusters (Figure 6.2). Some of these
clusters became progressively encrusted by nanospheres of silica (<250 nm in diameter) and eventually recolonised by a new succession of microbes
(Figure 6.3-6.4). Note that individual nanospheres
attached directly to the cells and the associated
EPS, thereby starting the process of silicification.
Cells that appeared free from these silica spheres
may be tentatively regarded as unsilicified.
The continuous interplay between microcolony formation, silicification, and recolonisation
appears to have formed vertically upright, pillar-like
microspicules (Figure 7). The upper surfaces of
these spicules displayed a knobby morphology
(typically 1-2 m in diameter) of similar dimensions
to nearby coccoidal cells (Figure 7.3). Thus we
infer that individual knobs formed by the silicification of these microbial cells. The knobby texture
was only observed on spicules of Microfacies 1,
where it coincided with the presence of coccoidal
microbial cells. We did not observe this texture on
spicules from other microfacies where bacilli (rodshapes) predominate (see below).
The irregularly lobed cocci are regarded as
microbial cells and not as abiogenic crystals, since
they are highly irregular in shape, often associated
with copious EPS (Figure 6) and are very similar in
morphology under SEM to cultured archaeal thermoacidophiles (e.g., Chen et al. 2005, figure 5A).
Indeed, 16S rRNA genes extracted from Microfacies 1 sinter is related to known thermoacidophiles
that also have irregularly lobed sphere shapes
(e.g., Thermoplasmatales; Schinteie 2005).

SCHINTEIE, CAMPBELL, & BROWNE: STROMATOLITIC MICROFACIES

1
2

4
5

N
8

9
10
12

11

Microfacies 2

13
14
6

19

Microfacies 1

Microfacies 4

21
22

Key:

25

Microfacies 1

24

Microfacies 2
Microfacies 3
Microfacies 4

Scale:
0

5m

Locality number (see below)

Figure 4. Characteristics of discharged thermal water from each microfacies. 4.1 Map of the study site showing
approximate microfacies locations. Numbers refer to localities were measurements were taken (see Figure 4.3). 4.2
Representative examples of glass slides with subaerially accreted silica after 67 days. Differences between slides
indicate different rates of subaerial silica accretion. Dashed blue line marks average water level. Note the relatively
large subaqueous sulphur accumulation in Microfacies 1. A yellow-light brown substance (sulphur? and/or organics?)
accumulated subaqueously in Microfacies 2. Bar represents ~3 cm. Silica accretion rates were unable to be measured for Microfacies 3.

SEM examination revealed that silica deposited predominantly as spheres of opal-A that are
primarily <10 nm to 50 nm in diameter. However,
spheres can grow up to 500 nm in diameter
through the aggregation of smaller spheres. Two
types of sphere shapes were observed: (1) freshly
deposited, round equidimensional spheres; and (2)
more poorly defined spheres with interparticle
necks (Figure 8.1-8.2) (cf. Iler 1979, figure 3.24).
The poorly defined sphere shape-texture was pervasive on Microfacies 1 sinter. Associations
between this texture and microorganisms were not
observed under SEM. In addition to silica, well
developed crystals of gypsum, barite, and sulphur
were present. No clay minerals were observed
8

anywhere upon or within sinter from this or any


other microfacies at the Parariki site. The absence
of clay minerals is in contrast to acid-derived sinters and residues reported elsewhere in New
Zealand (Jones et al. 2000; Rodgers et al. 2002).
At the microscale, the surfaces of the lower
portions of Microfacies 1 sinter are irregular and
uneven. Anastomosing ridges, nodules, and associated cavities are common. Incipient ridges
appear as surface irregularities that are bounded
by small cavities (Figure 8.3-8.4). Within these cavities, nodules formed that eventually become
mushroom-shaped, producing a neck-like structure
(Figure 8.4-8.6). While incipient nodules are sur-

PALAEO-ELECTRONICA.ORG

Figure 4 (continued) 4.3 Temperature, pH, water level changes and silica accretion measurements. N.R. = not
recorded. No significant water temperature changes were observed between seasons.

Microfacies
Stream

Water temperature
(summer/winter)
(C)
30.8/17

Water level (summer/


winter)
(mm)
150/159

49

55

58

57.9

8/12

2.23

N.R.

55

12/18

2.23

0.1289

54.6

12/15

N.R.

N.R.

58

12/16

2.3

0.1281

2 and 3

59.8

6/10

2.13

N.R.

10

45

No change

2.24

0.0092
0.0802

Locality
number
1

Water pH
2.03

Silica accumulation
over 67 days (g)
N.R.

12/14

2.3

0.1287

N.R.

2.25

N.R.

20/23

2.14

N.R.

11

85

10/12

2.10

12

64.2

10/14

2.13

N.R.

13

54

10/19

2.12

N.R.

14

38.8

130/136

N.R.

N.R.

15

63

5/8

2.24

N.R.

16

63.9

6/10

N.R.

N.R.

17

83.2

N.R.

2.28

0.2752

18

76.1

25/28

N.R.

N.R.

19

78.2

20/25

N.R.

0.2036
0.0098

20

67

No change

N.R.

21

84.6

N.R.

2.26

N.R.

22

59.4

18/20

2.23

N.R.

23

78.1

21/25

2.27

N.R.

24

52

No change

2.22

N.R.

25

91

N.R.

N.R.

N.R.

26

Stream

33.2/20.9

120/130

N.R.

N.R.

rounded by small cavities, taller nodules are associated with deeper cavities and larger ridges.
The lower portions of Microfacies 1 sinters
exhibit isolated patches of surficial silica sheets
(Figure 9.1). These sheets appeared to have
become progressively contracted and diminished,
particularly around necks (Figure 9.2), so as to
leave behind isolated remnants of a formerly continuous surface. The patches are bound by anastomosing ridges (Figure 9.3) that are larger towards
the outer fringes of these isolated patches (Figure
9.4). Figure 9.1-9.4 show the surface of silica
grown onto a glass slide for one month in a Microfacies 1 setting. Surfaces of natural Microfacies 1
sinter show even more extreme forms of isolated
remnants (Figure 9.5-9.6), presumably because
they have been in this environment for much longer
than the slides. These remnants contain concentric
laminae, mirroring the topography of the underlying
layers, as well as ridges, which are larger on the
fringes.

Microfacies 2 Spiculose Sinter


Microfacies 2 (Figure 3) is characterised by
quiescent thermal water discharge. Steam emission is also minor, while water temperatures range
from ~85C to ~30C (Figure 4.1). As in Microfacies 1, neither splash nor spray was observed
around these deposits. Microfacies 2 sinter forms
subaerially on pumice clasts, wood, pine cones,
and dead insects (Figure 10). Water level changes
have also been observed in this microfacies (Figure 4.1). Steam condensate keeps the sinters
moist. Silica accretion rates are slower than those
in Microfacies 1; an average of 0.12 g silica
accreted on the slides over 67 days (Figure 4).
The outer and lower portions of Microfacies 2
sinters are usually covered by a ubiquitous green
microbial mat (Figure 10). Transmission electron
microscopy (TEM) and rRNA gene analysis indicates that these mats are primarily composed of
coccoidal algal cells related to the rhodophyte
taxon Cyanidiophyceae (Cyanidium and Galdieria;
9

SCHINTEIE, CAMPBELL, & BROWNE: STROMATOLITIC MICROFACIES

Figure 5. Field occurrences and hand specimens of Microfacies 1 sinter. (1) Overview of Microfacies 1 represented
by a sulphur-emitting vents (84C) around which subaerial sinter forms (red arrows). White-grey and orange deposits
to the left and right belong to Microfacies 2 and 4, respectively. Green microbial mats form where water cools to
<52.5C. Note copious steam emission around sulphur discharging vent. Bar represents ~1.5 m. (2) Cup-shaped sinter (grey rim) depositing on the surface of a pumice clast. Spicules formed within a cavity on the uppermost side of the
rim (red arrow). Bar represents 3 cm. (3) Ridge-shaped sinter on a pumice clast. Bar represents 3.5 cm. (4) Close-up
of the lower portion of a Microfacies 1 sinter deposit. Multiple flat siliceous layers (yellow arrows) exhibit an irregular,
gnarled surface texture. Isolated patches of silica (red arrows) also occur. (5) Vertical section through sinter and pumice substratum. Note the alternating light brown and light grey laminae. (6) Thin-section micrograph showing microspicules (red arrows) along the length of a brown lamina. Laminae become increasingly more convex upwards.

Schinteie 2005). The mats occur where temperatures are between 52.5 and 30C. On sinter, the
mats are at their thickest (1 mm) at ~ 45C. On the
uppermost portions of the sinters, where temperatures are often slightly cooler and less exposed to
thermal water, the mats become faint and disappear.
The sinters are white to vitreous in colour, and
typified by significant spicular growth (Figure 10).
The spicules are needle-like to conical in shape, 1
10

to 3 mm wide, and from <1 mm to ~1 cm long. Spicules are densely packed, accrete perpendicular to
the substratum, and become progressively smaller
outward towards the water. A rim forms where spicules progressively link laterally with each other
through a continuous deposition of silica (Figure
10.2 and 10.5; cf. Handley 2004). Vertical sections
through these sinters revealed alternating light and
dark laminae (Figure 10.6).

PALAEO-ELECTRONICA.ORG

Figure 6. Microbial colonisation and microcolony formation associated with Microfacies 1 sinter. (1) Irregularly lobed,
coccoidal cells, and associated EPS. (2) Close-up of a conical microcolony. (3) Preferential silicification on the uppermost portion of a microcolony. (4) Colonisation of a silica substratum (yellow arrow) by a succession of partially silicified biofilm. Red arrows indicate acicular, aluminum, and oxygen-rich minerals (as determined by EDS) of unknown
phase.

The laminae alternate among laterally continuous green, brown, and translucent silica (Figure
11.1). As in Microfacies 1, laminae in the lower sinter portions are relatively flat but become progressively convex upwards (Figure 11.2), culminating in
spicules composed of parabolic laminae. The fate
of incipient spicules varied over time; some
became reinforced by the deposition of successive
laminae, whereas others were smothered or dampened by the lamination process (Figure 11.3 and
11.4). Overall, the relief of the sinters increased
over time, with surfaces characterised by abundant, erect, and free-standing spicular structures
with parabolic laminae. Numerous spicules also
exhibit branching and small (~0.5 mm) projections
with internal convex laminae.
The green laminae observed in thin section
(Figure 11.5) are composed of spheres that are
similar in size (2-10 m) and appearance to the
coccoidal cells present in the green biomats (Figure 11.6). These cells are restricted to the lower
and middle portions of Microfacies 2 sinter, corresponding to the distribution of the living green
mats. Diatom tests occur throughout the sinter.
As in Microfacies 1, sinter morphology also is
controlled by substrate shape and dimensions.
Subaerial substrate portions that are relatively

wider and higher above water exhibit less siliceous


encrustation relative to substrate size than deposits that are low-lying and have smaller widths.
While wider substrates display sinter with cupshaped morphologies (Figure 10.2), smaller substrates are covered in spicules that coat much of
the subaerial portions (Figure 10.3). The extent of
siliceous covering in Microfacies 2 is less than that
of comparable substrate sizes and shapes in
Microfacies 1, where relatively higher energy conditions occur.
Bacilli (i.e., rod-shaped microorganisms) (12.3 m long) dominate the sinter biota of this
microfacies (Figure 12.1). The microbes are usually associated with a meshwork of EPS that
includes both fibrous and mucosal textures. These
biofilms became progressively silicified by coalesced to partially coalesced opal-A nanospheres
(100 nm) and microspheres (250 nm) (Figure
12.2). Eventually, the films were completely obliterated by the deposition of overlying spheres.
The bacilli also tended to form clumps of
microcolonies which grew perpendicular to the silica surface and were associated with a network of
EPS (Figure 12.3). In places, diatom tests provided
areas of positive relief onto which new clusters of
bacilli aggregated. As in Microfacies 1, such micro11

SCHINTEIE, CAMPBELL, & BROWNE: STROMATOLITIC MICROFACIES

Figure 7. Formation of microspicular structures on Microfacies 1 sinter. (1) Upright microspicules at various stages of
growth and covered with microbial cells. (2) Cluster of cells attached to the top of a microspicule. (3) Knobby texture
on the uppermost portion of a microspicule with individual knobs (red arrows) equal in size to unsilicified cells (yellow
arrows).

colonies became silicified, forming areas of positive relief (Figure 12.4). This consistent interplay
between bacterial colonisation and silicification
also resulted in the formation of microspicules. The
possible outlines of bacilli inside spicules are preserved (Figure 12.6). The occurrence of vertically
upright microcolonies in Microfacies 2 is more
widespread than in Microfacies 1.
Pennate diatoms, predominantly Pinnularia
acoricola Hustedt and P. champmaniana Foged,
constitute a major component of the sinter biota
(Schinteie 2005; cf. Foged 1979; Cassie 1989;
Cassie and Cooper 1989). These diatoms preferentially occupied areas of low microrelief, such as
crevices and cracks, as well as overlying areas of
positive relief (Figure 13.1-13.3). The mode of
attachment for these benthic diatoms is adnate, or
closely appressed to the substratum, with the
entire valve attached to a substrate by a coating of
EPS (Figure 13.4). In open areas of the deposits
that do not offer sheltering by surrounding sinter,
12

diatom tests were often found fractured and


amassed into clumps (Figure 13.5). Diatom assemblages may eventually become part of the sinter
deposit, starting with the precipitation of silica
spheres onto the tests, which eventually result in
their complete cementation (Figure 13.6) (cf. Jones
et al. 2000; Campbell et al. 2004). In this and in
other microfacies of the Parariki site, early silicification occurs preferentially on the edges of diatom
tests, EPS sheets, and fibres.
The green algal-dominated mats present on
the lower surfaces and in thin section of Microfacies 2 sinters are largely composed of colonies of
spherical cells (2-10 m in diameter) that are covered in membranous sheets (Figure 14.1). The cell
surfaces of these mats can become encrusted and
incorporated into lower portions of the sinter (Figure 14.2-14.4; cf. Figure 11.5-11.6). Vertical sections through these portions exhibit numerous
horizontal laminae (~100-180 m thick) of dense,
vitreous sinter, alternating with layers (~80-120 m

PALAEO-ELECTRONICA.ORG

Figure 8. Formation of irregular silica surface textures on Microfacies 1 sinter. (1) Clusters of coalesced silica
spheres. (2) Close-up of (1) showing ill-defined outlines. (3) Anastomosing ridges and cavities at various stages of formation. (4) Incipient occurrence of ridges, associated cavities, and nodules (red arrows). Note broad deposition of
secondarily deposited silica spheres (yellow arrows). (5) Early-stage occurrence of nodules, acquiring a cone to round
shaped surface. (6) Late-stage occurrences of nodules, exhibiting necking (red arrows).

thick) dominated by silicified cells that are equal in


size to cells of the living green mats (Figure 14.5).
Several silicified cells also exhibited endospores
(i.e., internal division of parental algal cells). In
some instances, cellular impressions were preserved (Figure 14.6). The boundary between the
predominantly abiotic and biotic layers is sharp and
planar, showing that cells of the green mats initially
colonised a flat surface before becoming silicified
(Figure 14.4).
Unlike sinter surfaces of Microfacies 1, those
from Microfacies 2-4 are generally not irregular in
appearance. Ridges, nodules, isolated remnants,
or associated cavities were rarely observed. Furthermore, their opal-A spheres (also <10 nm to 50

nm in diameter) tended to be more spherical in


shape.
Microfacies 3 Parallel Laminated Sinter
Microfacies 3 sinter (Figure 3) is confined to
the southern portion of the study site, which slopes
at ~10 E from the horizontal. Although these
deposits are usually submerged under flowing thermal water (Figure 15), the water level in this area is
lower (~6-10 mm) than that flowing on flatter areas
(>10 mm) elsewhere at the site (Figure 4.1). Therefore, this microfacies is affected by changing water
levels, completely exposing the sinters to the air in
times of lower water levels. Water temperatures
range from ~60-54C (Figure 4.1).

13

SCHINTEIE, CAMPBELL, & BROWNE: STROMATOLITIC MICROFACIES

Figure 9. Isolated patches on the lower portions of Microfacies 1 sinter. (1) Overview of a sinter surface grown on a
glass slide for one month and showing isolated patches of surficial layers. (2) Necking of isolated patches (red arrows).
(3) Isolated patches exhibiting ridges around their sides. (4) Close-up of (3) showing ridges becoming progressively
smaller inwards towards the centre of the isolated patch (red arrow). (5) Isolated remnants exhibiting concentric laminae. (6) Remnant island exhibiting necking (yellow arrow) and numerous ridges (red arrows) that increase in size
towards the outer fringes.

Sinter from this environment is flat and parallel-laminated (Figure 15.1). Surfaces are irregular
and even rippled in places, which appears to be
due to the patchy nature of ongoing silica deposition (Figure 15.2). During lower water levels, small,
isolated puddles of thermal water occur in the irregular surface crevices of the sinters. Continuous
patchy silica deposition on these sinters results in
silica rims (Figure 15.3) that eventually grow into
cup-like deposits like those of Microfacies 2 (Figure
15.4). Indeed, pumice clasts that rest partially submerged in these waters and on top of the planar
14

sinter deposits act as substrates for Microfacies 2


sinters (Figure 15.1). Due to the thinness of Microfacies 3 and 4 sinter (~<3 mm and 2 mm, respectively), no thin sections were made of these two
deposit types. Because of difficulties in permanently placing glass slides horizontally, silica accretion rate measurements were also not conducted
for this microfacies. Any silica that deposits on vertically oriented slides would have represented
Microfacies 2 conditions.
SEM revealed that the upper surfaces of
Microfacies 3 sinters are colonised by bacilli (1-8

PALAEO-ELECTRONICA.ORG

Figure 10. Field occurrences and hand specimens of Microfacies 2 sinter. (1) Overview of Microfacies 2 with sinters
(white deposits) surrounded by relatively quiescent thermal water. A green microbial mat (algal) covers portions that
are 52.5C. Photo width is ~5 m. (2) Cup-shaped siliceous sinter (white rim and green/white spicules) depositing on
the upper surface of a pumice clast. Bar represents 4 cm. (3) Spiculose deposit covering entire substrate surface. Bar
represents 1.5 cm. (4) Cup-shaped and spiculose deposits forming in areas where thermal water flow is confined to
small spaces between densely clustered substrates. Coin at the lower right is 2 cm wide. Bar represents 2.5 cm. (5)
Side-on view of sinter showing a rim along inner portions of the deposit, while the outer portions are covered by vertical spicules. (6) Vertical section through sinter showing a white deposit with fine laminae. Sinter rests on a silicified
sandy substratum (brown). Bar represents 1 cm.

m long) (Figure 16.1). Silicification of these


microbes was patchy, with nanospheres (<100 nm)
of silica precipitating onto the uppermost portions
of cells, whereas those beneath were unsilicified
(Figure 16.2). Vertical sections of the sinters reveal
silica with a predominantly massive, vitreous tex-

ture (Figure 16.3). Cavities, cemented diatoms,


and spherical cells (2-10 m in diameter), resembling those from the green living mats, are the only
discernable features in these sections (Figure
16.3-16.4).

15

SCHINTEIE, CAMPBELL, & BROWNE: STROMATOLITIC MICROFACIES

2 mm

2 mm

1 mm

2 mm

100 m

Figure 11. Thin-section micrographs of Microfacies 2 sinter. (1) Branching spicules exhibiting projections (red arrow)
on their sides. (2) Close-up of sinter laminae that become increasingly convex upwards. (3) Reinforcement (red arrow)
or dampening (green arrow) of microspicules by the deposition of successive laminae. (4) Conical microspicules (red
arrows) growing on a lamina (cf. Figure 5.6). New layers of silica (green arrow) have subsequently covered the microspicules. (5) Overview of sinter displaying alternating laminae on the lower portions that are light brown and dark
green in colour (green arrow). (6) Close-up of green laminae present on the sinter surface (red arrow) and incorporated into the sinter body (green arrow).

On vertical portions of the sinters that have


subaerial rims (Figure 15.3-15.4), diatoms clustered together in large groups with the tips of their
tests aligned perpendicular to the approximate airwater interface (Figure 16.5). These clusters also
became progressively silicified and incorporated
into the deposit (Figure 16.6).
16

Microfacies 4 Thin Siliceous Rim Sinter


Sinters from Microfacies 4 (Figure 3) are characterised by thin (~2 mm thick), orange, cup-like
rims formed on small (<2 cm in diameter) pumice
clasts that rest upon moist sandy substrates well
above the outflow channels (Figure 17.1-17.3). The
deposits can reach heights of 1-6 mm, with rare or

PALAEO-ELECTRONICA.ORG

Figure 12. Microbial colonisation and microcolony formation on Microfacies 2 sinter. (1) Bacilli associated with
mucosal EPS. (2) Silicified mucosal EPS. (3) Conical microcolony formed predominantly by bacilli and fibrous EPS. (4)
and (5) Formation of a conical microspicule, showing the close association between the presence of biofilms and silicification. Red arrows in (4) indicate diatom shells. (6) Cross-section through a microspicule, showing it to be composed
of a network of rod-shapes (red arrows) that resemble bacilli. Yellow arrow points to a modern, unsilicified bacillus.

absent spicular textures. Vertical sections through


the rims reveal thin internal laminae (~0.5 mm
thick) that are convex rather than horizontal (Figure
17.4). The outer sides of these sinters are typically
covered with green algal mats. Rarely, anastomosing ridges occur on the inner sides of the sinter
rims (Figure 17.4).
Thermal water was not observed to wash,
splash, or spray onto these deposits. Small holes
dug into the sandy substrates revealed warm thermal pore water seeping upward through the sands
(Figure 17.3). This water may derive from surrounding thermal discharges nearby, or from small
unidentified vents underneath the sandy alluvium.
The temperature of the seeping water was

recorded to range from 45-67C (Figure 4.1). Silica


accretion rates in this microfacies averaged only
9.5 x 10-3 g over 67 days (Figure 4).
Sinter surfaces are covered by biofilms of diatoms (Figure 18.1-18.2), spherical microorganisms
(2-6 m in diameter) (Figure 18.3-18.4), and bacilli
(1-2.3 m long) (Figure 18.5). It is likely that the
spherical cells are also algae, belonging to the
green mats that cover the sinter rims on the margins of the deposits. Microbial silicification is likewise patchy, with some microbes completely
covered in silica spheres, while others nearby are
unsilicified. The bacilli also had the tendency to
form vertically upright microcolonies, but were
rarely observed to be silicified (Figure 18.5).
17

SCHINTEIE, CAMPBELL, & BROWNE: STROMATOLITIC MICROFACIES

Figure 13. Lifestyle modes of diatom assemblages on Microfacies 2 sinter. (1) Clusters of diatoms occupying numerous portions of a spicule. (2) Cluster of diatoms occupying a crevice on a spicule surface. (3) Diatoms wedged along
layers of silica that act as surfaces of positive relief. Note the succession of silica layers that drape a spicule surface
(red arrows), forming convex-upward layers similar to those observed in thin section (Figure 11.1). (4) Diatoms closely
appressed to the surface by a coating of mucosal EPS. (5) Fractured diatom cells amassed together in an open area
lacking shelter by surrounding sinter. (6) Cemented diatom cells incorporated into sinter.

SEM showed that vertical sections of these


sinters comprise a massive, vitreous texture with
no visible laminae. Silicified spherical cells (2-6 m
in diameter), often at the endospore stage, were
commonly incorporated into the sinter (Figure
18.6). However, these organisms do not form layers as in Microfacies 2, but are scattered across
the sinter.
DISCUSSION
This study assesses the effects of environmental variables on siliceous sinter deposition in
the acidic waters of the Parariki study site. The
18

restriction of different types of sinter morphologies,


textures and associated microorganisms with
respect to different microfacies shows that local
environmental factors are of profound influence.
Furthermore, a dynamic interplay between abiotic
and biotic factors plays a major role in forming sinter textures. In the following discussion we evaluate the significance of different environmental
variables (both abiotic and biotic) on the genesis of
the Parariki sinters. Where appropriate, we also
draw comparison with previous studies of acidderived siliceous hot spring deposits.

PALAEO-ELECTRONICA.ORG

Figure 14. Biotic composition and silicification of green algal-dominated mats on Microfacies 2 sinter. (1) Overview
of an unsilicified green mat showing a predominance of spherical cells that are 2-4 m in diameter. Minor amounts of
bacilli are also present. (2) Silicified portion of a green mat on the top (young) sinter surface. Note the incorporation of
equivalent sized spherical cells into the (older) sinter body (red arrows) (cf. Figure 11.6). (3) Alternating laminae of
sinter composed of silicified biomats (red arrows) and layers of featureless, massive, vitreous silica (yellow arrows).
(4) Sharp boundary (red arrow) between a biotic and an abiotic layer. (5) Close-up of silicified cells incorporated into
sinter beneath living green mat. Some silicified cells include endospores. (6) Preservation of cellular impressions (red
arrows).

Subaerial Silica Precipitation and


the Role of Microorganisms
Abiotic precipitation of silica was shown previously to occur where silica-rich waters wet subaerial substrates, resulting in increased silica
oversaturation due to cooling and evaporation
(e.g., Weed 1889; Krauskopf 1956; Walter et al.
1972; Rimstidt and Cole 1983; Hinman and Lindstrom 1996; Renaut et al. 1998; Campbell et al.
2002; Mountain et al. 2003). In the Parariki outflows, monomeric silica polymerises and deposits
opal-A as the water cools and evaporates (Teece
2000). However, the acid pH at the site slows silica
from precipitating subaqueously by inhibiting

monomeric silicic acid from deprotonating, polymerising, and nucleating (cf. Iler 1979; Makrides et
al. 1980; Weres et al. 1981; Mountain et al. 2003).
This inhibition explains both the monomeric form of
dissolved silica in the Parariki outflows and the lack
of subaqueous silica deposition. The low propensity for silicic acid to polymerise in acid environments is responsible for the small size (<100-500
nm) of the opal-A spheres that comprise the
Parariki sinters (cf. Iler 1979). However, monomeric silica may also deposit directly onto the subaerial substratum. Direct monomeric deposition,
concomitant with precipitation of small silica
spheres, can result in the formation of the massive
19

SCHINTEIE, CAMPBELL, & BROWNE: STROMATOLITIC MICROFACIES

Figure 15. Field occurrences and morphologies of Microfacies 3 sinter. (1) Flat, sinter deposit (brown colour) over
which thermal water flows (blue arrows). Microfacies 2 sinter forms on pumice clasts above water level (black arrows).
Hammer to the upper right is ~30 cm long. (2) Close-up of Microfacies 3 sinter showing irregular surface texture. (3)
Patchy silicification (yellow arrows) and the beginnings of rim formation (red arrow). The rims are frequently above
water level. Bar represents 2 cm. (4) Formation of a cup-like deposit (red arrow) that is continuous with the flat, parallel
laminated sinter below. Coin is 2.5 cm wide.

vitreous silica texture observed in the cross-sections of all sinters examined in this study (cf. White
et al. 1956; Rimstidt and Cole 1983; Handley et al.
2005).
While diatoms take up silicic acid from the
water to form their tests, prokaryotes are not
known to actively precipitate silica by metabolic
activity (Mountain et al. 2003; Konhauser et al.
2004). In this study, it appears that the silicification
of microbes and their associated EPS is a passive
process. Silicification often does not proceed
evenly on the biofilms of Parariki sinters, but is
concentrated in places that are more prone to cooling and evaporation, such as the upper portions of
microcolonies and spicules (Figure 6) (cf. Jones
and Renaut 1997; Lowe and Braunstein 2003).
Recent experiments have shown that silica precipitation is affected by its concentration but is independent of microbial growth (Toporski et al. 2002;
Yee et al. 2003; Benning et al. 2004). Nevertheless, microbial cell surfaces at the Parariki site
20

seem to act as favourable nucleation sites for silica


precipitation and polymerisation. Silica shows an
affinity with functional groups on proteins and
polysaccharides of cell walls and EPS (e.g.,
Schultze-Lam et al. 1995; Westall et al. 1995; Konhauser and Ferris 1996; Renaut et al. 1998;
Farmer 1999; Asada and Tazaki 2001; Konhauser
et al. 2001). However, microbial silicification in
acidic conditions may differ from those in near-neutral to alkaline waters. Acidophilic microbes, for
example, may act as reactive interfaces that promote silica nucleation and enhance precipitation
kinetics (Fortin and Beveridge 1997). Furthermore,
the cell walls of some acidophiles are considered
to adsorb hydrogen ions to their surfaces, forming
a static barrier against proton (H+) influx into cell
interiors (Gimmler et al. 1989; Asada and Tazaki
2001). Asada and Tazaki (2001) suggested that
highly reactive silica ions could be generated in
acidic hot springs when adsorbed hydrogen ions
combine with monomeric silicic acid, promoting sil-

PALAEO-ELECTRONICA.ORG

1
A

B
2

3
C

4
D

E5

6
F

Figure 16. Biotic characteristics and textures of Microfacies 3 sinter. (1) Colonisation of sinter surface by bacilli. (2)
Patchy silicification of cells by nanospheres (<100 nm) of silica. (3) Overview of sinter section showing a predominantly massive, vitreous texture, although cavities are also present. Silicified, sandy sediment underlies sinter. (4)
Diatoms and spherical cells cemented into sinter. (5) Clusters of cemented pennate diatoms aligned perpendicular to
the approximate air-water interface (dashed blue line). (6) Close-up of (5) showing aligned pennate diatoms.

ica polymerisation if a steady supply of silicic acid


is provided.
Microbial silicification at the Parariki site is
also important for the textural development of
these sinters. Mountain et al. (2003) noted that the
small silica sphere sizes occurring in low pH waters
can produce a dense silica matrix that accurately
preserves microbial matter. In this study, small
sphere sizes resulted in the intricate preservation
of microbial cell outlines and their EPS (Figures 6,
7 and 12). The precipitation of opal-A spheres on
diatom tests and their eventual cementation at the
Parariki site shows that these tests can act as sites
upon which silica precipitates (cf. Campbell et al.
2004; Jones et al. 2000). Indeed, microbial silicification occurs on all morphotypes at the Parariki

site, although species-specific patterns of silicification have been noted elsewhere (e.g., Francis et al.
1978; Westall et al. 1995; Toporski et al. 2002;
Lalonde et al. 2005). The tendency of silica to
deposit on the margins of diatoms and EPS (Figure
18.2) is likely due to the higher energy encountered
at these surfaces (Banfield and Hamers 1997).
Atoms on edges have lower coordination and
strongly asymmetric bonding configurations (Banfield and Hamers 1997), thereby allowing preferential deposition of silica. The patchy nature of
microbial silicification on Parariki sinters (Figure
16.2) may be induced abiotically, or by continuous
microbial growth and cell division. The latter would
enable microbial populations to survive during constant bathing by silica-rich thermal outflow.
21

SCHINTEIE, CAMPBELL, & BROWNE: STROMATOLITIC MICROFACIES

Figure 17. Field occurrences and hand specimens of Microfacies 4 sinter. (1) Thin siliceous sinter rims (orange colour, red arrow) forming on small pumice clasts that rest on fine sandy substrates. Coin in the centre of photo is 2 cm
wide. Bar represents ~15 cm. (2) Close-up showing underlying sandy substrate (covered by green microbial mat) that
is protected from erosion by the overlying orange sinter, thereby forming pillar-like structures. Bar represents 4 cm. (3)
Small hole dug into the sandy substrate revealing warm (45C) thermal pore water at shallow depths. Shovel to the
upper right is 15 cm wide. (4) Vertical section through a thin siliceous rim sinter. Pumice clast underneath sinter is covered with green microbial mat. Red arrow indicates thin internal laminae that are convex in shape. Bar represents 0.8
cm.

Subaerial Constraints on
Sinter Dimensions and Morphogenesis
The formation mechanisms and morphology
of Parariki sinters are primarily the products of substrate shape and size and the specific environment. Silica-rich water reaches subaerial
substrates in the outflows through water level
changes, wave wash and likely through capillary
creep. However, these mechanisms are microfacies dependent, and do not occur everywhere at
the study site.
In Microfacies 1-3, fluctuating water levels
enable thermal outflows to reach subaerial portions
of various substrates. Subsequent cooling and
evaporation ensue, allowing dissolved silica to
become oversaturated and deposit as continuous
layers. Wave wash, caused by the pulsating dis22

charge of thermal water, produces a similar effect.


However, this mechanism is largely confined to
Microfacies 1, where relatively vigorous vent discharge occurs.
Substrate width and shape governs the extent
of siliceous covering. On substrates that are relatively wide and higher above water, silica reaches
only the margins, producing a ring-like structure.
Continuous accumulation of silica between spicules can form a rim (cf. Handley 2004), which, if
high enough, blocks further silica deposition
inwards. Smaller substrates (e.g., pumice pebbles), in turn, allow silica-rich water to reach surfaces evenly, resulting in sinters that cover the
entire original surface. A similar process was suggested to form thrombolites at Lake Clifton, Western Australia (Moore and Burne 1994), and micro-

PALAEO-ELECTRONICA.ORG

Figure 18. Biotic characteristics and textures of Microfacies 4 sinter. (1) Clusters of interconnected diatoms (red
arrows) distributed on the surface of the sinter rim. Yellow arrow points to gypsum. (2) Close-up of interconnected diatoms (numbered), with silica spheres (~250 nm) precipitating on test margins (red arrow). (3) Colonisation of sinter
surface by spherical cells. (4) Spherical cell with silica precipitating on its surface (red arrow). (5) Vertically upright
microcolony of bacilli on a diatom test. (6) Incorporation of spherical cells, including endospores, into sinter.

atolls in modern corals (Stoddard and Scoffin


1979). In these two cases, upward growth is constrained by local water level, and subaerial exposure of the upper surfaces restricts growth to the
margins (Stoddard and Scoffin 1979; Moore and
Burne 1994).
The local energy of the thermal waters surrounding substrates can also affect sinter morphology. In Microfacies 1, where relatively more
turbulent conditions prevail, silica-rich water
reaches greater subaerial portions of the substrates compared to area- and shape-equivalent
deposits in Microfacies 2. Hence, sinters from the

former exhibit a greater extent of siliceous coating


than those from Microfacies 2 and 4, where more
quiescent conditions prevail. In places where thermal water availability is low due to an increasingly
sandy substratum, only the margins of substrates
become silicified. The resulting deposits are
observed in parts of Microfacies 2 (Figure 10.4)
and, especially, Microfacies 4 (Figure 17), where
thermal water moves around increasingly finer
grain sizes between adjacent sinter substrate
deposits.
Capillary creep and/or diffusion may also play
a role in subaerial sinter formation (cf. Henley
23

SCHINTEIE, CAMPBELL, & BROWNE: STROMATOLITIC MICROFACIES

1996; Hinman and Lindstrom 1996; Renaut et al.


1998, 1999; Campbell et al. 2002; Guidry and
Chafetz 2002; Mountain et al. 2003), particularly for
the cup-shaped deposits of Microfacies 4. Substrates there are neither exposed to changing
water levels, nor are they affected by wave wash.
Thermal pore water that is present within the sandy
substrate is therefore most likely drawn up towards
the surface by capillary rise or diffusion. An upward
rise of thermal water would explain the often nearvertical rims of these deposits and their confinement to only the small pumice substrate margins.
In this scenario, the margins form convex-upward
laminae (Figure 17.4). Small, calcareous stromatolitic structures of similarly low relief (<5 cm in
diameter) occur on a sandy shoreline at Lake Clifton, Western Australia (Moore and Burne 1994).
Even when exposed by low lake levels during summer, these stromatolites remain saturated by low
salinity groundwater seepage through capillary
action (Moore and Burne 1994).
In Microfacies 3, silica precipitation will be
slowed by the acidic discharge that flows over the
flat, parallel-laminated deposits (Figure 15.1). However, periodic exposure above the water and concomitant cooling and evaporation of remaining
puddles of silica-rich water would allow for sinter to
accumulate. The patchy silicification of these sinters (Figure 15.3) is likely a result of cooling and
evaporation of local puddles of thermal water.
However, other factors not identified in this study
may also play a role in the formation of Microfacies
3 sinter.
Significance of Microbial Biofilms upon Sinter
Biofilms can afford protection from environmental extremes (Hall-Stoodley et al. 2004), which
are common in an acid hot spring setting: low pH
(McNeill and Hamilton 2003); metal toxicity (Teitzel
and Parsek 2003); dehydration and high salinity
(Le Magrex-Debar et al. 2000; Sutherland 2001);
and UV exposure (Espeland and Wetzel 2001). In
addition, associated EPS may prevent cell silicification by providing reactive sites for silica to bind
(Lalonde et al. 2005). According to contemporary
models (e.g., Stoodley et al. 2002), the formation
and development of prokaryotic biofilms requires
the transport of microbes to a surface and their initial attachment, followed by microcolony formation.
In quiescent waters of low-shear, laminar flow, and
with ideal nutrient conditions, microcolonies often
resemble pillar, mushroom, or mound-like structures (e.g., Hall-Stoodley and Stoodley 2002;
Stoodley et al. 2002; Hall-Stoodley et al. 2004;
24

Purevdorj et al. 2002). These morphologies are


formed by clonal division, whereby daughter cells
spread outwards and upwards from the attachment
surface to form cell clusters (Hall-Stoodley and
Stoodley 2002; Stoodley et al. 2002). Microcolonies on the Parariki sinters also exhibit positive
relief, with both bacilliform and coccoidal prokaryotic microorganisms developing vertically upright
structures. At the Parariki site, lobed coccoidal and
bacilli morphotypes also were confined to different
microfacies settings, with the former largely
restricted to Microfacies 1, and the latter observed
elsewhere at the site. Similar differential distributions of microbial morphotypes occur in acid hot
springs at Yellowstone (e.g., Brock 1978), Italy
(Simmons and Norris 2002), and Montserrat (Burton and Norris 2000). In these studies, lobed coccoidal
prokaryotic
microorganisms
(e.g.,
Sulfolobus) occur in higher temperature settings
(usually >60C), while bacilli (e.g., Thiobacillus) are
found in the relatively cooler waters (>30C).
Apart from prokaryotes, algae, particularly diatoms and members of the Cyanidiophyceae, are
ubiquitous in the relatively cooler waters (52.5C)
at the Parariki site. The predominance of algal
mats at the study site is consistent with previous
studies of acidic hot springs in New Zealand (Brock
and Brock 1971; Brock 1978; Cassie and Cooper
1989; Jones et al. 2000) and elsewhere (e.g.,
Brock 1973, 1978; Gross 1998; Seckbach 1998;
Ferris et al. 2005; Walker et al. 2005). The role of
Cyanidiophyceae-dominated biofilms in the formation of laminae is discussed below. Diatom biofilms
also are an important constituent of sinters from
Microfacies 2-4, where waters are cool enough for
diatom survival (typically 45C; Cassie 1989).
Benthic diatoms that are adnate, or closely
appressed to the substratum, like those at the
Parariki site, tend to be motile (Cohn and Dispari
1994; Cohn and Weitzell 1996). Such diatoms
glide up through sediments in a movement that is
non-random, following distinctive sets of chemotactic and phototactic responses (Cohn and Disparti
1994; Cohn and Weitzell 1996). EPS secretion by
these organisms is also used for their daily migrations across surfaces (Cohn and Weitzell 1996).
Diatoms that attach to siliceous deposits at
the Parariki site prefer to inhabit areas of low
microrelief, such as pits, cracks, and along small
cavities (Figure 13.1-13.3). In studies conducted
elsewhere, protective areas were shown to provide
refugia for diatoms from grazers (e.g., Dudley and
DAntonio 1991), shield them from abrasion and
drag associated with moving water (Luttenton and

PALAEO-ELECTRONICA.ORG

Rada 1986), and protect them from desiccation


(Hostetter and Hoshaw 1970). The effects of grazing activity on diatoms at the Parariki site are
unknown. However, the presence of numerous
clumps of fractured diatom shells indicates that
water is turbulent enough in places to crush diatom
tests and transport them (Figure 13.5). Repeated
wetting and drying of the siliceous deposits could
also cause significant stress to diatoms.
While a preponderance of diatoms has been
noted in other acid hot spring settings (Jones et al.
2000), no fungi were observed on Parariki sinters.
This observation is in contrast to previous studies
of acid hot spring deposits, where fungi are purported to be dominant (Jones et al. 1999, 2000).
Origins of Spiculose Textures
While spicules from the Parariki sinters are
similar in their gross morphology to spicular geyserite around spouters (cf. Walter 1976; Braunstein
and Lowe 2001; Jones and Renaut 2003; Lowe
and Braunstein 2003), there are also distinctive differences between them. Spicular geyserite forms
at the air-water interface where water splashes
around the inner rims of springs (Walter 1976).
While biofilms are present on spicular geyserite
and help form porous laminae (Cady and Farmer
1996), the formation of these spicules is largely
attributed to abiotic mechanisms, involving splash
and spray of silica-rich waters (Walter 1976; Braunstein and Lowe 2001; Jones and Renaut 2003;
Lowe and Braunstein 2003). Spicules that form this
way tend to be larger at the poolward sides of rims,
where greater wave activity and splash occurs,
than on the landward sides (Walter 1976; Lowe
and Braunstein 2003). At the Parariki site, by contrast, neither splash nor spray was observed, and
spicules are progressively larger away from the
water. Furthermore, the Parariki spicules are bigger where water is more quiescent, with longer spicules in Microfacies 2 (1 cm in high), than in the
more turbulent Microfacies 1 (0.5 cm high) environment. In the latter, spicules occur only on the
uppermost portions of the sinters, away from the
thermal waters, and within cavities (Figure 5.2).
Thus, the spiculose Parariki sinter textures must
develop by mechanisms different from those of
classic spicular geyserite.
In previous studies, biotic mechanisms for
spicular sinter formation have also been proposed
(Cassie and Cooper 1989; Campbell et al. 2002)
and shown to be facilitated by filamentous
microbes at Champagne Pool, New Zealand (Handley et al. 2005). In this study, microspicule initia-

tion and development was observed to involve a


dynamic interplay between biofilm growth and silica deposition. Vertically upright colonies of microorganisms (coccoidal and bacilliform) were
observed to act as domains of positive relief that
commonly became progressively silicified and subsequently recolonised (cf. Handley et al. 2005).
The restriction of these microspicules to quiescent
sinter portions correlates with the distribution of
their macroscopic counterparts.
It was beyond the scope of this study to establish causes governing the development of the
Parariki biofilm morphologies. However, biofilm
development is a multifactorial process influenced
by both environmental factors and genes (HallStoodley et al. 2004). Previous studies have shown
how environmental conditions, such as water flow,
including turbulence, can potentially affect biofilm
development, superseding cell-cell communications as a principal determinant of biofilm morphogenesis (Hall-Stoodley and Stoodley 2002;
Purevdorj et al. 2002). The tendency for vertically
upright microcolonies to form in less turbulent
waters may therefore explain the restriction of the
Parariki spicules to more quiescent areas. Nevertheless, environmental factors other than water turbulence, such as nutrient availability and
phototrophy, may also play a role (e.g., Doemel
and Brock 1974; Brock 1978; Stoodley et al. 2002;
Hall-Stoodley et al. 2004). However, for spicule formation to proceed and continue, the supply of silica
is likewise important. In Microfacies 4, thermal
water is confined to pore spaces within the sandy
substrate. Water is not turbulent there and sinters
experience only minor contact with thermal water.
Therefore, sinter growth rates are slowest in that
setting and spicules are generally absent, although
vertically upright microcolonies are present.
Hence, balances between environmentally induced
biofilm morphogenesis and the supply of silica-rich
water are most likely the major determinants of spicule formation and growth at the Parariki site (cf.
Handley et al. 2005). Fluctuations in the supply of
silica would reinforce or dampen spicule growth by
the deposition of successive silica laminae (Figure
11.3). Such fluctuations could be achieved by dilution of the thermal water by rain water or diurnal or
seasonal variations in temperature.
The occurrence and preservation of spicular
textures in fossil sinter deposits may be of importance to palaeoenvironmental analysis. However,
care should be applied when interpreting textures.
As noted above, spiculose sinters can occur in
both quiescent and in turbulent conditions, the lat25

SCHINTEIE, CAMPBELL, & BROWNE: STROMATOLITIC MICROFACIES

ter producing geyserite. Therefore, spiculose textures in ancient hot spring deposits should not be
taken as the sole facies indicator of water turbulence but considered in context with other proxies
(e.g., oncoids and pisoids; silicified streamers; preserved sinter terraces) (e.g., Walter et al. 1996;
Campbell et al. 2001).
Morphogenesis of Ridges, Cavities,
Nodules, and Isolated Remnants
Anastomosing ridges, associated cavities,
nodules, and pits that occur on surfaces of Microfacies 1 sinter were seen previously on siliceous sinter (Braunstein and Lowe 2001; Lowe and
Braunstein 2003) and silica residue (Cook et al.
2000; Rodgers et al. 2002, 2004). For sinter, constructive processes were inferred for their formation. Such constructive processes result when pits
retain water between wetting events from geyser
eruptions, while capillary action and evaporation
draw water along edges, forming rims alongside
these pits through concomitant deposition of silica
(Lowe and Braunstein 2003, figure 20A). In crosssection, these deposits are typically composed of
cavities and pseudo-cross-laminae that mark pit
migration (Lowe and Braunstein 2003, figure 20B,
C). Micropitted nodular sinters, or knobs, are also
suggested to form from repeated wetting and drying (Braunstein and Lowe 2001, figure 12B).
The formation of ridges, cavities, and nodules
on Microfacies 1 sinter at the Parariki site may
involve a similar component of construction. However, destructive processes in acid settings should
not be discounted. Ridge formation does not necessarily require repeated wetting events by silicarich water, as similar irregular ridges have been
observed to form in sinter buried in humate-rich soil
that was exposed to rain water (B.Y. Lynne, personal commun., 2004). Nodules, similar in shape
to those in the Parariki sinters of Microfacies 1,
also occur on siliceous coverings of pumice on
steaming ground at Rotokawa (R. Schinteie,
unpublished data).
Anastomosing ridges on silica residue are
interpreted to be destructive remnants of surfaces
etched by dissolution (Cook et al. 2000; Rodgers et
al. 2002, 2004). Alongside these ridges, layers of
once-cemented opal-A microspheres become
exposed, while depressions between the ridges
are lined by irregular clusters of silica spheres at
different stages of dissolution (cf. Rodgers et al.
2002, figure 8). Corrosive activity could likewise
produce the gnarly texture (Figure 5.4), necking
(Figures 8.6, 9.2 and 9.6), and the dominance of
26

larger ridges on the outermost sides of isolated


remnants (Figure 9.4 and 9.6) on Microfacies 1 sinter. Corrosive attack focused around edges, kink
sites, or necks would likely occur due to the lower
coordination number and hence lower energy
encountered in these locations (cf. Banfield and
Hamers 1997).
Cross-sections of Parariki sinters, seen in thin
section and under SEM, reveal that they are massive in texture, with neither open cavities nor
pseudo-cross-laminae present. Therefore, the
ridge and cavity morphology is only a surface feature. A later-stage deposition of silica would bury
these irregular surface textures.
Steam condensate, made acid by the oxidation of H2S, is often cited as the cause of corrosion
of silica residue or sinter surfaces close to vents or
steaming ground (e.g., Rodgers et al. 2002; Jones
and Renaut 2003; Lynne and Campbell 2004).
Complexing of silica by sulphate has been suggested to increase amorphous silica solubilities in
aqueous Na2SO4 solutions (Marshall and Chen
1982; Fournier and Marshall 1983). Acid steam
condensate may therefore explain the restriction of
corrosive-like textures to predominantly Microfacies 1 sinter where copious steam emission
occurs. However, further research is warranted.
Changes in Silica Sphere Morphology
The ill-defined silica-spheres (Figure 8.1-8.2)
in Microfacies 1 sinter superficially appear as if
they were covered by microbially induced, mucuslike substances. However, the pervasive nature of
this texture and the lack of association with microorganisms may call for alternative explanations.
Similarly ill-defined sphere shapes have been
observed on acid-formed sinters elsewhere and
attributed to repeated episodes of silica dissolution
and redeposition, causing a blurring of particle
detail (Rodgers et al. 2004). Silica solubility is
greater on convex surfaces than on concave surfaces (Iler 1979). Hence, silica spheres become
obscure in form as silica dissolves from the upper
convex surfaces and redeposits on the concave
surfaces, where solubility is lower. This process
forms interparticle necks. Rodgers et al. (2004)
suggested that changes in microchemical conditions were responsible for the episodic nature of
silica dissolution and redeposition in acid water
derived sinters. Similarly, near fumaroles, indistinct
opal-A spheres form closely packed aggregates
(Lynne and Campbell 2004). In silica residue,
gelatinous, ill-defined spheres (also referred to as
frog spawn texture) form and are purported to be

PALAEO-ELECTRONICA.ORG

the result of progressive strengthening of interparticle bonds at the contact between adjacent silica
spheres (Rodgers et al. 2002, 2004).
Origins of Sinter Laminae
While laminae in Parariki sinters of all microfacies were visible in hand sample and/or in thin section, they were not observed under SEM, where
vertical sections mostly revealed a massive, vitreous sinter body (e.g., Figure 16.3). However, the
irregular, gnarled surfaces on Microfacies 1 sinter
exhibited layering under SEM (Figure 9.5), while
for Microfacies 2 deposits, a succession of silica
layers was observed only at the surface (Figure
13.3). Jones and Renault (2004) related indistinctive laminae in other sinters to differences in the
water contents of opal-A. Since these differences
are texturally featureless, they are not detected by
standard SEM methods. However, etching of these
surfaces, attributed to acidic steam condensate,
reveals alternating laminae caused by differential
dissolution of opal-A due to local differences in silica solubility (Jones and Renaut 2003, 2004).
Jones and Renaut (2004) interpreted wet opal
layers to be formed by rapid precipitation of
hydrated silica. Dry opal, in turn, would form by
slow evaporation, resulting in layers of silica with
less water. Since the Parariki sinters also form by
repeated wetting and drying, a continuous laminated buildup would likewise develop.
Lower portions of sinter from Microfacies 2, by
contrast, consist of laminae that are both abiotic
and biotic in nature. Green, algal-dominated mats
cover a large portion of these sinters close to the
air-water interface (Figure 10), and become silicified and incorporated into the sinter (Figure 14).
Studies of the taxonomically related alga Cyanidium caldarium, suggest that silicification of its mats
result by the continuous proliferation of non-motile
algal cells, with parental cells underlying younger
cells (Asada and Tazaki 1999). Therefore, older
cells will face shortages of light and CO2 for photosynthesis, and O2 for respiration. Asada and
Tazaki (1999) suggested that these stresses impair
the ability of the algal cells to regulate silica on their
walls, so that silica continuously grows on them
and progressively fills the interstices between silica
crusts of different cells.
The restriction of the green Parariki algal mats
to the lower portions of Microfacies 2 sinters (Figures 10 and 11.5) could be temperature controlled,
although moisture and nutrient supply may also
influence their distributions. Brock (1978) found C.
caldarium cells growing in high densities at ~45C

throughout its range (55-20C). Indeed, the green


mats from our study are thickest (1 mm) close to
the water level at Microfacies 2 (50-40C), and
gradually disappear closer to the upper portions of
the sinters. On the higher substrate areas of Microfacies 2 sinters, where green mats are absent, silica layering occurs only as fine laminae in thinsection and as a succession of silica horizons
under SEM (Figure 13.3, red arrows), with no
apparent microbial association. Continuous deposition of silica by wetting and drying appears to be
the principal method of forming these abiotic laminae in the upper areas.
Stromatolitic Nature of Parariki Sinters
The sinters forming at the Parariki study site
are laminated growth structures, morphologically
similar in appearance to stromatolites described
around other hot springs. However, the criteria
used to define stromatolites are not straightforward
because the search for a clear and widely
accepted definition of them has proven controversial (Ginsburg 1991; Cady et al. 2003). Generally,
most authors define stromatolites as layered organosedimentary structures, formed by the trapping,
binding, and/or precipitation of sediments as a
result of the growth and metabolic activity of microorganisms (Walter 1976; Krumbein 1983). However, the lack of fossilised microbes in many
ancient stromatolites, or the potential for stromatolitic laminae to be formed entirely by abiotic
means (Grotzinger and Rothman 1996) has
caused problems with this genetic definition.
At the Parariki site, sinter laminae and textures such as spicules formed as a result of a combination of abiotic and biotic factors. However,
taphonomic constraints can affect the differential
preservation of microorganisms. At the Parariki
site, ubiquitous microbial mats of prokaryote-sized
microorganisms cover sinters from all microfacies.
Nevertheless, preservation of these microbes as
distinct, silicified laminae is often absent or confined to spicules. By contrast, algal cells, including
diatoms, are much better preserved in Parariki sinter laminae. This differential preservation would
bias the potential fossil record in assuming that
these sinters were once largely colonised by
eukaryotes. Therefore, a descriptive and nongenetic definition of stromatolites has been
adopted herein, whereby a stromatolite is
described as an attached, laminated, lithified sedimentary growth structure, accretionary away from
a point of limited surface of initiation (Semikhatov
et al., 1979).
27

SCHINTEIE, CAMPBELL, & BROWNE: STROMATOLITIC MICROFACIES

CONCLUSIONS

REFERENCES

Environmental variables, acting on a microscale over a few centimetres (or less) at the
Parariki study site, have a differential effect on sinter textures and microbiology. Hence, four different
microfacies of stromatolitic sinter occur, featuring
distinctive deposit morphologies, textures, formation mechanisms, and microbial communities. Constraints on their distribution include: 1) the size and
shape of the subaerial substrate surfaces; 2) the
relative exposure to thermal waters; 3) water flow
rate; 4) the presence, nature, and taphonomy of
mat-forming microorganisms; 5) water and ambient
temperature; and 6) potential exposure to acid
steam condensate. The dynamic interplay between
silica deposition and microbial activity is a major
factor in the formation of a variety of sinter textures, particularly spicules. Laminae develop in
these sinters due to abiotic and/or alternating abiotic and biotic processes. Different microbial communities also exhibit variable preservation potential
as sinter accumulates, regardless of microbial
abundances.
Results of this study are consistent with previous studies of acid hot spring deposits. In terms of
the associated biota, eukaryotes, namely diatoms
and the rhodophyte taxon Cyanidiophyceae, are a
ubiquitous sinter component at temperatures
<52.5C. Lobed coccoidal and bacilliform morphotypes of prokaryotic dimensions also were confined
to different microfacies settings, with the former
largely restricted to the hotter setting of Microfacies
1, and the latter observed elsewhere at the site.
Clay minerals, as well as fungi, were not observed
in this study. Mineralogically, sinters of Microfacies
1 are similar to silica residue and siliceous deposits
around fumaroles; namely a gnarled, possibly corrosive, surface texture and/or faint silica sphere
shapes.

Asada, R. and Tazaki, K. 1999. Biomineralization of silica


under strong acidic conditions. Proceedings of the
International Symposium, Kanazawa, Earth-WaterHumans, 209-216.
Asada, R. and Tazaki, K. 2001. Silica biomineralization
of unicellular microbes under strongly acidic conditions. The Canadian Mineralogist, 39:1-16.
Banfield, J.F. and Hamers, R.J. 1997. Processes at minerals and surfaces with relevance to microorganisms
and prebiotic synthesis, p. 81-117. In Banfield, J.F.,
and Nealson, K.H. (eds.), Geomicrobiology: interactions between Microbes and Minerals. Mineralogical
Society of America, Washington D.C.
Barns, S.M., Delwiche, C.F.D., Palmer, J.D., Dawson,
S.C., Hershberger, K.L., and Pace, N.R. 1996. Phylogenetic perspective on microbial life in hydrothermal
ecosystems, past and present, p. 24-39. In Bock,
G.R., and Goode, J.A. (eds.), Evolution of Hydrothermal Ecosystems on Earth (and Mars?), Ciba Foundation Symposium 202. John Wiley and Sons,
Chichester.
Benison, K.C. and Laclair, D.A. 2002. Acid sedimentary
environments on Mars?: possible terrestrial analogs.
The Geological Society of America, Denver Annual
Meeting, 34:174.
Benning, L.G., Phoenix, V.R., Yee, N., and Konhauser,
K.O. 2004. The dynamics of cyanobacterial silicification: an infrared micro-spectroscopic investigation.
Geochimica et Cosmochimica Acta, 68:743-757.
Bibby, H.M., Caldwell, T.G., Davey, F.J., and Webb, T.H.
1995. Geophysical evidence on the structure of the
Taupo Volcanic Zone and its hydrothermal circulation. Journal of Volcanology and Geothermal
Research, 68:29-58.
Blank, C.E., Cady, S.L., and Pace, N.R. 2002. Microbial
composition of near-boiling silica-depositing thermal
springs throughout Yellowstone National Park.
Applied and Environmental Microbiology, 68:51235135.
Braunstein, D. and Lowe, D.R. 2001. Relationship
between spring and geyser activity and the deposition and morphology of high temperature (>73C) siliceous sinter, Yellowstone National Park, Wyoming,
U.S.A. Journal of Sedimentary Research, 71:747763.
Brock, T.D. 1973. Lower pH limit for the existence of
blue-green algae: evolutionary and ecological implications. Science, 179:480-483.
Brock, T.D. 1978. Thermophilic Microorganisms and Life
at High Temperatures. Springer Verlag, New York.
Brock, T.D. and Brock, M.L. 1971. Microbiological studies of thermal habitats of the central volcanic region,
North Island, New Zealand. New Zealand Journal of
Marine and Freshwater Research, 5:233-258.
Browne, P.R.L. 1974. Rotokaua geothermal field. Minerals of New Zealand, New Zealand Geological Survey,
Report No. 38.

ACKNOWLEDGEMENTS
We thank J. Tahau (Tauhara Trust) and the
New Zealand Department of Conservation for
granting us site access. Funding support for water
chemistry analyses was supplied by the Marsden
Fund, Royal Society of New Zealand. Technical
assistance was provided by A. Arcila, L. Cotterall,
C. Hobbis, B. James, R. Sims, A. Turner, J. Wilmshurst, and B. Wong. Insightful discussions were
held with K. Brown, K. Handley, N. Hinman, M.
Hochstein, B. Lynne, B. Mountain, B. Ricketts and
S. Turner. We also thank two anonymous reviewers for their valued comments.
28

PALAEO-ELECTRONICA.ORG

Bullock, M.A. 2005. The flow and ebb of water. Nature,


438:1087-1088.
Burton, N.P. and Norris, P.R. 2000. Microbiology of
acidic, geothermal springs of Montserrat: environmental rDNA analysis. Extremophiles, 4:315-320.
Cady, S.L. and Farmer, J.D. 1996. Fossilization processes in siliceous thermal springs: trends in preservation along thermal gradients, p. 150-173. In Bock,
G.R. and Goode, J.A. (eds.), Evolution of Hydrothermal Ecosystems on Earth (and Mars?), Ciba Foundation Symposium 202. John Wiley and Sons,
Chichester.
Cady, S.L., Farmer, J.D., Grotzinger, J.P., Schopf, W.J.,
and Steele, A. 2003. Morphological biosignatures
and the search for life on Mars. Astrobiology, 3:351368.
Campbell, K.A., Buddle, T.F., and Browne, P.R.L. 2004.
Late Pleistocene siliceous sinter associated with fluvial, lacustrine, volcaniclastic and landslide deposits
at Tahunaatara, Taupo Volcanic Zone, New Zealand.
Transactions of the Royal Society of Edinburgh:
Earth Sciences, 94:485-501.
Campbell, K.A., Rodgers, K.A., Brotheridge, J.M.A., and
Browne, P.R.L. 2002. An unusual modern silica-carbonate sinter from Pavlova spring, Ngatamariki, New
Zealand. Sedimentology, 49:835-854.
Campbell, K.A., Sannazzaro, K., Rodgers, K.A., Herdianita, N.R., and Browne, P.R.L. 2001. Sedimentary
facies and mineralogy of the late Pleistocene
Umukuri silica sinter, Taupo Volcanic Zone, New
Zealand. Journal of Sedimentary Research, 71:727746.
Cassie, V. 1989. A contribution to the study of New
Zealand diatoms. Bibliotheca Phycologica, Band 17,
J. Cramer, Berlin, Stuttgart.
Cassie, V. and Cooper, R.C. 1989. Algae of New
Zealand thermal areas. Bibliotheca Phycologica,
Band 78, J. Cramer, Berlin, Stuttgart.
Chen, Z.-W., Jiang, C.-Y., She, Q., Liu, S.-J., and Zhou,
P.-J. 2005. Key role of cysteine residues in catalysis
and subcellular localization of sulfur oxygenasereductase of Acidianus tengchongensis. Applied and
Environmental Microbiology, 71:621-628.
Cohn, S.A. and Disparti, N.C. 1994. Environmental factors influencing diatom cell motility. Journal of Phycology, 30:818-828.
Cohn, S.A. and Weitzell, Jr., R.E. 1996. Ecological considerations of diatom cell motility. I. Characterization
of motility and adhesion in four diatom species. Journal of Phycology, 32:928-939.
Collar, R.J. and Browne, P.R.L. 1985. Hydrothermal
eruptions in the Rotokawa Geothermal System,
Taupo Volcanic Zone, New Zealand. Proceedings of
the 7th Geothermal Workshop, Geothermal Institute,
University of Auckland, Auckland, 171-176.

Cook, K.L., Rodgers K.A., Campbell K.A., Browne,


P.R.L., Martin, R., and Seakins, J.M. 2000. The mineralogy, texture and significance of silica residue
from the Te Kopia geothermal field, Taupo Volcanic
Zone. Proceedings of the 22nd New Zealand Geothermal Workshop, Geothermal Institute, University
of Auckland, Auckland, 143-149.
Doemel, W.N. and Brock, T.D. 1974. Bacterial stromatolites: origin of laminations. Science, 184:1083-1085.
Dove, P.M. and Rimstidt, J.D. 1994. Silica-water interactions, p. 259-308. In Heaney, P.J., Prewitt, C.T., and
Gibbs, G.V. (eds), Silica: Physical Behavior,
Geochemistry and Materials Applications. Mineralogical Society of America, Washington, D.C.
Dudley, T.L. and DAntonio, C.M. 1991. The effects of
substrate texture, grazing, and disturbance on macroalgal establishment in streams. Ecology, 72:297309.
Ellis, A.J. and Wilson, S.H. 1961. Hot spring areas with
acid-sulphate-chloride waters. Nature, 191:696-697.
Espeland, E.M. and Wetzel, R.G. 2001. Complexation,
stabilization, and UV photolysis of extracellular and
surface-bound glucosidase and alkaline phosphotase: implications for biofilm microbiota. Microbial
Ecology, 42:572-585.
Farmer, J.D. 1996. Hydrothermal systems on Mars: an
assessment of present evidence, p. 273-299. In
Bock, G.R. and Goode, J.A. (eds.), Evolution of
Hydrothermal Ecosystems on Earth (and Mars?),
Ciba Foundation Symposium 202. John Wiley and
Sons, Chichester.
Farmer, J.D. 1999. Taphonomic modes in microbial fossilization. Size Limits of Very Small Microorganisms:
Proceedings of a Workshop. Washington National
Academy Press, Washington, D.C.
Farmer, J.D. 2000. Hydrothermal systems: doorways to
early biosphere evolution. GSA Today, 10:2-9.
Farmer, J.D. and Des Marais, D.J. 1999. Exploring for a
record of ancient Martian life. Journal of Geophysical
Research, 104:26977-26995.
Ferris, M.J., Sheehan, K.B., Khl, M., Cooksey, K., Wigglesworth-Cooksey, B., Harvey R., and Henson, J.M.
2005. Algal species and light microenvironment in a
low pH, geothermal microbial mat community.
Applied and Environmental Microbiology, 71:71647171.
Foged, N. 1979. Diatoms in New Zealand, the North
Island. Bibliotheca Phycologica, Band 47, J. Cramer,
Berlin, Stuttgart.
Forsyth, D.J. 1977. Limnology of Lake Rotokawa and its
outlet stream. New Zealand Journal of Marine and
Freshwater Research, 11:525-539.
Fortin, D. and Beveridge, T.J. 1997. Role of the bacterium Thiobacillus in the formation of silicates in acidic
mine tailings. Chemical Geology, 141:235-250.
Fournier, R.O. 1985. The behavior of silica in hydrothermal solutions, p. 45-61. In Berger, B.R. and Bethke,
P.M. (eds.), Geology and Geochemistry of Epithermal Systems, Society of Economic Geologists.
29

SCHINTEIE, CAMPBELL, & BROWNE: STROMATOLITIC MICROFACIES


Fournier, R.O. and Marshall, W.L. 1983. Calculation of
amorphous silica solubilities at 25 to 300C and
apparent cation hydration numbers in aqueous salt
solutions using the concept of effective density of
water. Geochimica et Cosmochimica Acta, 47:587596.
Francis, S., Margulis, L., and Barghoorn, E.S. 1978. On
the experimental silicification of microorganisms II.
On the time of appearance of eukaryotic organisms
in the fossil record. Precambrian Research, 6:65100.
Gimmler, H., Weiss, U., and Weiss, C. 1989. pH-regulation and membrane potential of the extremely acid
resistant green alga Dunaliella acidophila, p. 389390. In Dainty, J. (ed.), Plant Membrane Transport.
Elsevier Science, Venice.
Ginsburg, R.N. 1991. Controversies about stromatolites:
vices and virtues, p. 25-36. In Mller, D.W., McKenzie, J.A., and Weissert, H. (eds.), Controversies in
Modern Geology. Academic Press Limited, London.
Grange, L.I. 1937. The geology of the Rotorua-Taupo
subdivision. Department of Scientific and Industrial
Research, Geological Survey Bulletin No. 37, Wellington.
Gross, W. 1998. Revision of comparative traits for the
acido- and thermophilic red algae Cyanidium and
Galdiera, p. 439-447. In Seckbach, J. (ed.), Enigmatic microorganisms and life in extreme environments. Kluwer Academic Publishers, Boston.
Grotzinger, J.P. and Rothman, D.H. 1996. An abiotic
model for stromatolite morphogenesis. Nature,
383:423-425.
Guidry, S.A. and Chafetz, H.S. 2002. Factors governing
subaqueous siliceous sinter precipitation in hot
springs: examples from Yellowstone National Park,
USA. Sedimentology, 49:1253-1267.
Guidry, S.A. and Chafetz, H.S. 2003. Anatomy of siliceous hot spring: examples from Yellowstone
National Park, Wyoming, USA. Sedimentary Geology, 157:71-106.
Hall-Stoodley, L. and Stoodley, P. 2002. Developmental
regulation of microbial biofilms. Current Opinion in
Biotechnology, 13:228-233.
Hall-Stoodley, L., Costerton, J.W., and Stoodley, P. 2004.
Bacterial biofilms: from the natural environment to
infectious diseases. Nature Reviews Microbiology,
2:95-108.
Handley, K.M. 2004. In situ experiments on the growth
and textural development of subaerial microstromatolites, Champagne Pool, Waiotapu, NZ. Unpublished
MSc Thesis, University of Auckland, Auckland.
Handley, K.M., Campbell, K.A., Mountain, B.W., and
Browne, P.R.L. 2005. Abiotic-biotic controls on the
origin and development of spicular sinter: in situ
growth experiments, Champagne Pool, Waiotapu,
New Zealand. Geobiology, 3:93-114.

30

Healy, J. 1965. Quaternary pumice deposits, p. 61-72. In


Thompson, B.N., Kermode, L.D., and Ewart, A.
(eds.), New Zealand Volcanology, Central Volcanic
Zone. Department of Scientific and Industrial
Research, Series 50, Wellington.
Henley, R.W. 1996. Chemical and physical context for
life in terrestrial hydrothermal systems: chemical
reactors for the early development of life and hydrothermal ecosystems, p. 61-82. In Bock, G.R. and
Goode, J.A. (eds.), Evolution of Hydrothermal Ecosystems on Earth (and Mars?), Ciba Foundation
Symposium 202. John Wiley and Sons, Chichester.
Herdianita, N.R., Rodgers, K.A., and Browne, P.R.L.
2000a. Routine instrumental procedures to characterise the mineralogy of modern and ancient sinter.
Geothermics, 29:65-81.
Herdianita, N.R., Browne, P.R.L., Rodgers, K.A., and
Campbell, K.A. 2000b. Mineralogical and textural
changes accompanying ageing of siliceous sinter.
Mineralium Deposita, 35:48-62.
Hinman, N.W. and Lindstrom, R.F. 1996. Seasonal
changes in silica deposition in hot spring systems.
Chemical Geology, 132:237-246.
Hostetter, H.P. and Hoshaw, R.W. 1970. Environmental
factors affecting resistance to desiccation in the diatom Stauroneis anceps. American Journal of Botany,
57:512-518.
Hugenholtz, P., Pitulle, C., Herschenberger, K.L., and
Pace, N.R. 1998. Novel division level bacterial diversity in a Yellowstone hot spring. Journal of Bacteriology, 180:366-376.
Iler, R.K. 1979. The chemistry of silica: solubility, polymerization, colloid and surface properties and biochemistry. Wiley, New York.
Inagaki, F., Motomura, Y., Doi, K., Taguchi, S., Izawa, E.,
Lowe, D.R., and Ogata, S. 2001. Silicified microbial
community from at Steep Cone Hot Spring, Yellowstone National Park. Microbes and Environment,
16:125-130.
Jones, B. and Renaut, R.W. 1997. Formation of silica
oncoids around geysers and hot prings at El Tatio,
northern Chile. Sedimentology, 44:287-304.
Jones, B. and Renaut, R.W. 2003. Petrography and genesis of spicular and columnar geyserite from the
Whakarewarewa and Orakeikorako geothermal
areas, North Island, New Zealand. Canadian Journal
of Earth Sciences, 40:1585-1610.
Jones, B. and Renaut, R.W. 2004. Water content of opalA: implications for the origin of laminae in geyserite
and sinter. Journal of Sedimentary Research, 74:117128.
Jones, B., Renaut, R.W., and Konhauser, K.O. 2005.
Genesis of large siliceous stromatolites at Frying Pan
Lake, Waimangu geothermal field, North Island, New
Zealand. Sedimentology, 52:1229-1252.
Jones, B., Renaut, R.W. and Rosen, M.R. 1999. Role of
fungi in the formation of siliceous coated grains,
Waiotapu geothermal field, North Island, New
Zealand. Palaios, 14:475-492.

PALAEO-ELECTRONICA.ORG

Jones, B., Renaut, R.W. and Rosen, M.R. 2000. Stromatolites forming in acidic hot-spring waters, North
Island, New Zealand. Palaios, 15:450-475.
Kerr, R.A. 2004. Rainbow of Martian minerals paints picture of degradation. Science, 305:770-771.
Konhauser, K.O. and Ferris, F.G. 1996. Diversity of iron
and silica precipitation by microbial mats in hydrothermal waters, Iceland: implications for Precambrian
iron formations. Geology, 24:323-326.
Konhauser, K.O., Jones, B., Phoenix, V.R., Ferris, G. and
Renaut, R.W. 2004. The microbial role in hot spring
silicification. Ambio, 33:552-558.
Konhauser, K.O., Phoenix, V.R., Bottrell, S.H., Adams,
D.G., and Head, I.M. 2001. Microbial-silica interactions in Icelandic hot spring sinter: possible analogues for some Precambrian siliceous stromatolites.
Sedimentology, 48:415-433.
Krauskopf, K.B. 1956. Dissolution and precipitation of silica at low temperatures. Geochimica et Cosmochimica Acta, 10:1-26.
Krumbein, W.E. 1983. Stromatolites: the challenge of a
term in time and space. Precambrian Research,
20:493-531.
Krupp, R.E., Browne, P.R.L., Henley, R.W., and Seward,
T.M. 1986. Rotokawa Geothermal Field, p. 47-55. In
Henley, R.W., Hedenquist, J.W., and Roberts, P.J.
(eds.), Guide to the Active Epithermal (Geothermal)
Systems and Precious Metal Deposits of New
Zealand, Monograph Series on Mineral Deposits,
Volume 26. Gebrder Borntraeger, Stuttgart.
Krupp, R.E. and Seward, T.M. 1987. The Rotokawa geothermal system, New Zealand. Economic Geology,
82:1109-1129.
Laclair, D.A. and Benison, K.C. 2002. Acid environments
on Mars?: the physical sedimentology. The Geological Society of America, Denver Annual Meeting,
34:175
Lalonde, S.V., Konhauser, K.O., Reysenbach, A.-L., and
Ferris, F.G. 2005. The experimental silicification of
Aquificales and their role in hot spring sinter formation. Geobiology, 3:41-52.
Le Magrex-Debar, E., Lemoine, J., Gelle, M.P., Jaqueline, L.F., and Choisy, C. 2000. Evaluation of biohazards in dehydrated biofilms on foodstuff packaging.
International Journal of Food Microbiology, 5:239243.
Lowe, D.R. and Braunstein, D. 2003. Microstructure of
high-temperature (>73C) siliceous sinter deposited
around hot springs and geysers, Yellowstone Park:
the role of biological and abiological processes in
sedimentation. Canadian Journal of Earth Sciences,
40:1611-1642.
Lowe, D.R., Anderson, K.S., and Braunstein, D., 2001.
The zonation and structuring of siliceous sinter
around hot springs, p. 143-166. In Reysenbach, A.L., Voytek, M., and Mancinelli, R. (eds.), Thermophiles: Biodiversity, Ecology and Evolution. Kluwer
Academic/Plenum Publishers, New York.

Luttenton, M.R. and Rada, R.G. 1986. Effects of disturbance on epiphytic community architecture. Journal
of Phycology, 22:320-326.
Lynne, B.Y. and Campbell, K.A. 2003. Diagenetic transformations (opal-A to quartz) of low- and mid-temperature microbial textures in siliceous hot-spring
deposits, Taupo Volcanic Zone, New Zealand. Canadian Journal of Earth Sciences, 40:1679-1696.
Lynne, B.Y. and Campbell, K.A. 2004. Morphologic and
mineralogic transitions from opal-A to opal-CT in lowtemperature siliceous sinter diagenesis, Taupo Volcanic Zone, New Zealand. Journal of Sedimentary
Research, 74:561-579.
Makrides, A.C., Turner, M., and Slaughter, J. 1980. Condensation of silica from supersaturated silicic acid
solutions. Journal of Colloid and Interface Science,
73:345-367.
Marshall, W.L. and Chen, C.-T.A. 1982. Amorphous silica solubilities VI. Postulated sulfate-silicic acid
solution complex. Geochimica et Cosmochimica
Acta, 46:367-370.
McCollom, T.M. and Hynek, B.M. 2005. A volcanic environment for bedrock diagenesis at Meridiani Planum
on Mars. Nature, 438:1129-1131.
McNeill, K. and Hamilton, I.R. 2003. Acid tolerance
response of biofilm cells of Streptococcus mutans.
FEMS Microbiology Letters, 221:25-30.
Moore, L.S. and Burne, R.V. 1994. The modern thrombolites of Lake Clifton, Western Australia, p. 3-29. In
Bertrand-Sarfati, J. and Monty, C. (eds.), Phanerozoic Stromatolites II. Kluwer Academic Publishers,
Boston.
Mountain, B.W., Benning, L.G., and Boerema, J.A. 2003.
Experimental studies on New Zealand hot spring sinters: rates of growth and textural development. Canadian Journal of Earth Sciences, 40:1643-1667.
Pace, N.R. 1997. A molecular view of microbial diversity
and the biosphere. Science, 276:734-740.
Pancost, R.D., Pressley, S., Coleman, J.M., Benning,
L.G., and Mountain B.W. 2005. Lipid biomolecules in
silica sinters: indicators of microbial biodiversity.
Environmental Microbiology, 7:66-77.
Pancost, R.D., Pressley S., Coleman, J.M., Talbot.,
H.M., Kelly, S.P., Farrimond, P., Schouten, S., Benning, L., and Mountain, B.W. 2006. Composition and
implications of diverse lipids in New Zealand geothermal sinters. Geobiology, 4:71-92.
Purevdorj, B., Costerton, J.W., and Stoodley, P. 2002.
Influence of hydrodynamics and cell signaling on the
structure and behavior of Pseudomonas aeruginosa
biofilms. Applied and Environmental Microbiology,
68:4457-4464.
Renaut, R.W., Jones, B., and Tiercelin, J.-J. 1998. Rapid
in situ silicification of microbes at Loburu hot springs,
Lake Bogoria, Kenya Rift Valley. Sedimentology,
45:1083-1103.

31

SCHINTEIE, CAMPBELL, & BROWNE: STROMATOLITIC MICROFACIES


Reyes, A.G., Trompetter, W.J., Britten, K., and Searle, J.
2002. Mineral deposits in the Rotokawa geothermal
pipelines, New Zealand. Journal of Volcanology and
Geothermal Research, 119:215-239.
Rice, C.M., Ashcroft, W.A., Batten, D.J., Boyce, A.J.,
Caulfield, J.B.D., Fallick, A.E., Hole, M.J., Jones, E.,
Pearson, M.J., Rodgers, G., Saxton, J.M., Stuart,
F.M., Trewin, N.H., and Turner, G. 1995. A Devonian
auriferous hot spring system, Rhynie, Scotland. Journal of the Geological Society, London, 152:229-250.
Rimstidt, J.D. and Cole, D.R. 1983. Geothermal mineralization I: the mechanism of formation of the
Beowawe, Nevada, siliceous sinter deposit. American Journal of Science, 283:861-875.
Rodgers, K.A., Browne, P.R.L., Buddle, T.F., Cook, K.L.,
Greatrex, R.A., Hampton, W.A., Herdianita, N.R.,
Holland, G.R., Lynne, B.Y., Martin, R., Newton, Z.,
Pastars, D., Sannazarro, K.L., and Teece, C.I.A.
2004. Silica phases in sinters and residues from geothermal fields of New Zealand. Earth Science
Reviews, 66:1-61.
Rodgers, K.A., Cook, K.L., Browne, P.R.L., and Campbell, K.A. 2002. The mineralogy, texture and significance of silica derived from alteration by steam
condensate in three New Zealand fields. Clay Minerals, 37:299-322.
Schinteie, R. 2005. Siliceous sinter facies and microbial
mats from acid-sulphate-chloride springs, Parariki
Stream, Rotokawa Geothermal Field, Taupo Volcanic
Zone, New Zealand. Unpublished MSc Thesis, University of Auckland, Auckland.
Schultze-Lam, S., Ferris, F.G., Konhauser, K.O., and
Wiese, R.G. 1995. In situ silicification of an Icelandic
hot spring microbial mat: implications for microfossil
formation. Canadian Journal of Earth Sciences,
32:2021-2026.
Seckbach, J. 1998. The cyanidiophyceae: hot spring acidophilic algae, p. 427-435. In Seckbach, J. (ed.),
Enigmatic microorganisms and life in extreme environments. Kluwer Academic Publishers, Boston.
Semikhatov, M.A., Gebelein, C.D., Cloud, P., Awramik,
S.M., and Benmore, W.C. 1979. Stromatolite morphogenesis: progress and problems. Canadian Journal of Earth Sciences, 19:992-1015.
Simmons, S. and Norris, P.R. 2002. Acidophiles of saline
water at thermal vents of Volcano, Italy. Extremophiles, 6:201-207.
Stetter, K.O. 1996. Hyperthermophiles in the history of
life, p. 1-10. In Bock, G.R. and Goode, J.A. (eds.),
Evolution of Hydrothermal Ecosystems on Earth
(and Mars?), Ciba Foundation Symposium 202. John
Wiley and Sons, Chichester.
Stoddart, D.R. and Scoffin, T.P. 1979. Microatolls: review
of form, origin and terminology. Atoll Research Bulletin, 224:1-17.
Stoodley, P., Sauer, K., Davies, D.G., and Costerton,
J.W. 2002. Biofilms as complex differentiated communities. Annual Review of Microbiology, 55:187209.
32

Sutherland, I.W. 2001. Biofilm exopolysaccharides: a


strong and sticky framework. Microbiology, 147:3-9.
Teece, C.I.A. 2000. Sinters deposited from acid-sulfate
chloride waters at the Rotokawa geothermal field
(Taupo Volcanic Zone, New Zealand). Unpublished
MSc Thesis, University of Auckland, Auckland.
Teitzel, G.M. and Parsek, M.R. 2003. Heavy metal resistance of biofilm and planktonic Pseudomonas aeruginosa. Applied and Environmental Microbiology,
61:2313-2320.
Toporski, J., Steele, A., Westall, F., Thomas-Keprta, K.L.,
and McKay, D.S. 2002. The simulated silicification of
bacteria new clues to the modes and timing of bacterial preservation and implications for the search for
extraterrestrial microfossils. Astrobiology, 2:1-26.
Trewin, N.H., Fayers, S.R., and Kelman, R. 2003. Subaqueous silicification of the contents of small ponds
in an Early Devonian hot-spring complex, Rhynie,
Scotland. Canadian Journal of Earth Sciences,
40:1697-1712.
Vucetich, C.G. and Pullar, W.A. 1973. Holocene tephra
formations erupted in the Taupo Area, and interbedded tephras from other volcanic sources. New
Zealand Journal of Geology and Geophysics,
16:745-780.
Walker, J.J., Spear, J.R., and Pace, N.R. 2005. Geobiology of a microbial endolithic community in the Yellowstone geothermal environment. Nature, 434:10111014.
Walter, M.R. 1976. Geyserites of Yellowstone National
Park: an example of abiogenic stromatolites, p. 87112. In Walter, M.R. (ed.), Stromatolites. Elsevier,
Amsterdam.
Walter, M.R., Des Marais, D., Farmer, J.D., and Hinman,
N.W. 1996. Lithofacies and biofacies of Mid-Paleozoic thermal spring deposits in the Drummond Basin,
Queensland, Australia. Palaios, 11:497-518.
Walter, M.R. and Des Marais, D.J. 1993. Preservation of
biological information in thermal spring deposits:
developing a strategy for the search for fossil life on
Mars. Icarus, 101:129-143.
Walter, M.R., Bauld, J., and Brock, T.D. 1972. Siliceous
algal and bacterial stromatolites in hot spring and
geyser effluents of Yellowstone National Park. Science, 178:402-405.
Weed, W.H. 1889. Formation of travertine and siliceous
sinter by the vegetation of hot springs. United States
Geological Survey, 9th Annual Report, 613-676.
Weres, O., Yee, A., and Tsao, L. 1981. Kinetics of silica
polymerization. Journal of Colloid and Interface Science, 84:379-402.
Westall, F., Boni, L., and Guerzoni, E. 1995. The experimental silicification of microorganisms. Palaeontology, 38:495-528.
White, D.E., Brannock, W.W., and Murata, K.J. 1956. Silica in hot-spring waters. Geochimica et Cosmochimica Acta, 10:27-59.

PALAEO-ELECTRONICA.ORG

Wilson, C.J.N., Houghton, B.F., McWilliams, M.O., Lanphere, M.A., Weaver, S.D., and Briggs, R.M. 1995.
Volcanic and structural evolution of Taupo Volcanic
Zone: a review. Journal of Volcanology and Geothermal Research, 68:1-28.

Yee, N., Phoenix, V.R., Konhauser, K.O., Benning, L.G.,


and Ferris, F.G. 2003. The effects of cyanobacteria
on silica precipitation at neutral pH: implications for
bacterial silicification in geothermal hot springs.
Chemical Geology, 199:83-90.

33

Vous aimerez peut-être aussi